id
stringlengths 27
33
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 13
1.81M
|
---|---|---|---|
warning/0006/hep-th0006034.html
|
ar5iv
|
text
|
# 1 Introduction.
## 1 Introduction.
One of the most remarkable properties of maximal ($`d=11`$) supergravity is the emergence of hidden symmetries of exceptional type in the reduction to lower dimensions . Some time ago, it was shown that, already in eleven dimensions, this theory permits reformulations where the tangent space symmetry $`SO(1,10)`$ is replaced by the local symmetries that would arise in the reduction to four and three dimensions, i.e. $`SO(1,3)\times SU(8)`$ and $`SO(1,2)\times SO(16)`$, respectively. The key point here is, of course, that this construction works without any assumptions restricting the dependence of theory on the coordinates, so these symmetries already exist in eleven dimensions. Besides its general interest, this result plays a pivotal role in establishing the consistency of the Kaluza Klein reduction of $`d=11`$ supergravity in various non-trivial backgrounds . In this article, we will concentrate on the $`SO(1,2)\times SO(16)`$ version of , which exhibits special features (especially โmaximal unificationโ of symmetries), and which in the reduction to three dimensions yields directly maximal $`N=16`$ supergravity<sup>1</sup><sup>1</sup>1The $`E_{8(8)}`$ invariance of maximal $`N=16`$ supergravity in three dimensions was originally shown in , while its complete Lagrangian and supersymmetry transformations were derived in . The latter follow directly by dimensional reduction of the variations presented in . See also for a more recent treatment..
The equivalence of the different versions of $`d=11`$ supergravity is established at the level of the equations of motion by making special gauge choices, and does not extend off shell because the new versions mix equations of motion and Bianchi identities of the original theory. As shown in , the bosonic fields, that would become scalar matter fields in the dimensional reduction, can be assigned to representations of the hidden global symmetry groups $`E_{7(7)}`$ and $`E_{8(8)}`$, respectively. For this purpose it was necessary to fuse the bosonic fields into new objects, christened โgeneralized vielbeineโ: these are soldering forms with upper world indices running over the internal dimensions, and lower indices belonging to the $`\mathrm{๐๐}`$ and $`\mathrm{๐๐๐}`$ representations of $`E_{7(7)}`$ and $`E_{8(8)}`$, respectively. The generalized vielbeine are subject to algebraic constraints which follow from their explicit expressions in terms of the vielbein of $`d=11`$ supergravity. For the $`SO(1,2)\times SO(16)`$ version of , there exists a differential constraint in addition, which is a $`d=11`$ variant of the duality constraint by which vector fields are converted into scalars in three dimensions. However, it was not clear from previous work how to solve these constraints (apart from the torus reduction, explicit solutions are only known for the $`S^7`$ compactification of $`d=11`$ supergravity ), and how to recover the correct counting of bosonic physical degrees of freedom.
An essential new element of the present work in comparison with previous results is our treatment of the 3-form potential $`A_{MNP}`$. Whereas in this field appeared only via its field strength $`F_{MNPQ}=24_{[M}A_{NPQ]}`$ inside the $`E_{7(7)}`$ and $`E_{8(8)}`$ connections given there, part of it here is merged into an enlarged generalized vielbein. As a consequence, the latter now also transforms under tensor gauge transformations, suggesting a partial unification of coordinate and tensor gauge transformations. We show that the generalized vielbein is actually part of a full $`E_{8(8)}`$ matrix $`๐ฑ`$ (a โ248-beinโ), which now lives in eleven dimensions, and which incorporates the bosonic degrees of freedom of $`d=11`$ supergravity, with the exception of the dreibein remaining from the 3+8 split (which does not propagate in the dimensionally reduced theory). Moreover, we present evidence for the existence of an โexceptional geometryโ for $`d=11`$ supergravity by displaying the action of the combined internal coordinate and tensor gauge transformations on this 248-bein. In deriving these results, we will make crucial use of certain special properties of the exceptional Lie algebra $`E_{8(8)}`$, in particular the existence of a maximal nilpotent abelian subalgebra of dimension 36, which is unique up to conjugation , and whose importance was recently emphasized in .
The present paper deals mainly with the algebraic relations obeyed by the generalized vielbein, and their solution. The corresponding differential relations will be discussed elsewhere. We believe that our results constitute further evidence for a hidden $`E_{8(8)}`$ structure of $`d=11`$ supergravity, but there remain a number of open problems that must be dealt with; these include in particular the proper treatment of the tensor components $`B_{\mu \nu m}`$ and $`B_{\mu \nu \rho }`$, and the construction of an invariant action in terms of the 248-bein $`๐ฑ`$ in eleven dimensions.
## 2 $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ invariant $`๐
\mathbf{=}\mathrm{๐๐}`$ supergravity
We will first review $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ formulation of $`๐
\mathbf{=}\mathrm{๐๐}`$ supergravity, referring readers to for further details. Our conventions concerning $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ as well as its $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ and $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ decompositions, which played an important role also in , are summarized in two appendices.
$`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ invariant $`๐
\mathbf{=}\mathrm{๐๐}`$ supergravity is derived from the original version of by first splitting up the fields in a way that would be appropriate for the reduction to three dimensions, but without dropping the dependence on any coordinates, and then reassembling the pieces into new objects transforming under local $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$. Hence we are still dealing with $`๐
\mathbf{=}\mathrm{๐๐}`$ supergravity, albeit in a very different guise. This is achieved by first breaking the original tangent space symmetry $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐๐}\mathbf{)}`$ down to $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ by a partial gauge choice for the elfbein, and then reenlarging it to $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ by the introduction of new gauge degrees of freedom. The construction thus requires a 3+8 split of the $`๐
\mathbf{=}\mathrm{๐๐}`$ coordinates and tensor indices. The main task then is to identify the proper $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ covariant fields and to verify that all supersymmetry variations as well as the equations of motion can be entirely expressed in terms of the new fields.
In a first step one thus brings the elfbein into triangular form by (partial) use of local $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐๐}\mathbf{)}`$ Lorentz invariance
$$๐ฌ_{๐ด}^{}{}_{}{}^{๐จ}\mathbf{=}\mathbf{\left(}\begin{array}{cc}๐ซ^\mathbf{}\mathrm{๐}๐_๐^{\underset{\mathbf{ยฏ}}{๐ถ}}& ๐ฉ_๐{}_{}{}^{๐}๐_{๐}^{}{}_{}{}^{๐}\\ \mathrm{๐}& ๐_{๐}^{}{}_{}{}^{๐}\end{array}\mathbf{\right)}\mathbf{,}๐ซ\mathbf{:=}๐๐๐ญ๐_๐{}_{}{}^{๐}\mathbf{,}$$
(2.1)
Here curved $`๐
\mathbf{=}\mathrm{๐๐}`$ indices decompose as $`๐ด\mathbf{=}\mathbf{(}๐\mathbf{,}๐\mathbf{)}\mathbf{,}๐ต\mathbf{=}\mathbf{(}๐\mathbf{,}๐\mathbf{)}\mathbf{,}\mathbf{}`$ with $`๐\mathbf{,}๐\mathbf{,}\mathbf{}\mathbf{=}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{,}\mathrm{๐}`$ and $`๐\mathbf{,}๐\mathbf{,}\mathbf{}\mathbf{=}\mathrm{๐}\mathbf{,}\mathbf{}\mathbf{,}\mathrm{๐๐}`$, and the associated flat indices are denoted by $`\underset{\mathbf{ยฏ}}{๐ถ}\mathbf{,}\underset{\mathbf{ยฏ}}{๐ท}\mathbf{,}\mathbf{}`$ and $`๐\mathbf{,}๐\mathbf{,}\mathbf{}`$, respectively<sup>2</sup><sup>2</sup>2We hope that no confusion results from our double usage of $`A,B,\mathrm{}`$ as both $`SO(1,10)`$ vector and $`SO(16)`$ spinor indices. It should always be clear from the context which is meant. (as in the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}`$ indices $`\underset{\mathbf{ยฏ}}{๐ถ}\mathbf{,}\underset{\mathbf{ยฏ}}{๐ท}\mathbf{,}\mathbf{}`$ are underlined to distinguish them from the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ spinor indices to be used below). The partially gauge fixed elfbein, whose form is preserved by the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ subgroup of $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐๐}\mathbf{)}`$, thus contains the Weyl rescaled dreibein $`๐_๐^๐ถ`$, the Kaluza-Klein vector $`๐ฉ_๐^๐`$ and the achtbein $`๐_{๐}^{}{}_{}{}^{๐}`$ yielding the scalar degrees of freedom living in the coset $`๐ฎ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}\mathbf{/}๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$.
The remaining bosonic degrees of freedom reside in the 3-index field $`๐จ_{๐ด๐ต๐ท}`$, which gives rise to various scalar and tensor fields upon performing a 3+8 split of the indices. First of all, there are 56 scalars $`๐จ_{๐๐๐}`$ and 28 vector fields
$$๐ฉ_{๐๐๐}\mathbf{:=}๐จ_{๐๐๐}\mathbf{}๐ฉ_{๐}^{}{}_{}{}^{๐}๐จ_{๐๐๐}$$
(2.2)
If one were to reduce to three dimensions, the 8+28 vector fields $`๐ฉ_๐^๐`$ and $`๐ฉ_{๐๐๐}`$ would be converted to 36 scalar degrees of freedom by means of a duality transformation. Here, they will be kept together with their dual scalars contained in the generalized vielbein, to which they are related by a nonlinear analog of the (linear) duality constraint of the reduced $`๐
\mathbf{=}\mathrm{๐}`$ supergravity.
In addition, $`๐จ_{๐ด๐ต๐ท}`$ gives rise to (always in the 3+8 split)
$`๐ฉ_{๐๐๐}`$ $`\mathbf{:=}`$ $`๐จ_{๐๐๐}\mathbf{}\mathrm{๐}๐ฉ_{\mathbf{[}๐}^{}{}_{}{}^{๐}๐จ_{๐\mathbf{]}๐๐}\mathbf{+}๐ฉ_{๐}^{}{}_{}{}^{๐}๐ฉ_{๐}^{}{}_{}{}^{๐}๐จ_{๐๐๐}`$
$`๐ฉ_{๐๐๐}`$ $`\mathbf{:=}`$ $`๐จ_{๐๐๐}\mathbf{}\mathrm{๐}๐ฉ_{\mathbf{[}๐}^{}{}_{}{}^{๐}๐จ_{๐๐\mathbf{]}๐}\mathbf{+}\mathrm{๐}๐ฉ_{\mathbf{[}๐}^{}{}_{}{}^{๐}๐ฉ_{๐}^{}{}_{}{}^{๐}๐จ_{๐\mathbf{]}๐๐}`$ (2.3)
$`\mathbf{}๐ฉ_{๐}^{}{}_{}{}^{๐}๐ฉ_{๐}^{}{}_{}{}^{๐}๐ฉ_{๐}^{}{}_{}{}^{๐}๐จ_{๐๐๐}`$
These fields are subject to the tensor gauge transformations
$$๐น๐จ_{๐ด๐ต๐ท}\mathbf{=}\mathrm{๐}\mathbf{}_{\mathbf{[}๐ด}๐_{๐ต๐ท\mathbf{]}}$$
(2.4)
Under these, we have
$`๐น๐ฉ_{๐๐๐}`$ $`\mathbf{=}`$ $`๐_๐๐_{๐๐}\mathbf{+}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐ฉ_{๐}^{}{}_{}{}^{๐}๐_{๐\mathbf{]}๐}\mathbf{+}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}\stackrel{\mathbf{~}}{๐}_{๐\mathbf{]}๐}`$
$`๐น๐ฉ_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathbf{}_๐\stackrel{\mathbf{~}}{๐}_{๐๐}\mathbf{+}\mathrm{๐}\mathbf{}_๐๐ฉ_{\mathbf{[}๐}^{}{}_{}{}^{๐}\stackrel{\mathbf{~}}{๐}_{๐๐\mathbf{]}}\mathbf{+}\mathrm{๐}๐_{\mathbf{[}๐}\stackrel{\mathbf{~}}{๐}_{๐\mathbf{]}๐}\mathbf{+}๐_{๐๐}^{}{}_{}{}^{๐}๐_{๐๐}`$
$`๐น๐ฉ_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathrm{๐}๐_{\mathbf{[}๐}\stackrel{\mathbf{~}}{๐}_{๐๐\mathbf{]}}\mathbf{}\mathrm{๐}๐_{\mathbf{[}๐๐}^{}{}_{}{}^{๐}\stackrel{\mathbf{~}}{๐}_{๐\mathbf{]}๐}`$ (2.5)
where
$$๐_๐\mathbf{:=}\mathbf{}_๐\mathbf{}๐ฉ_{๐}^{}{}_{}{}^{๐}\mathbf{}_๐$$
(2.6)
and
$$๐_{๐๐}^{}{}_{}{}^{๐}\mathbf{:=}๐_๐๐ฉ_{๐}^{}{}_{}{}^{๐}\mathbf{}๐_๐๐ฉ_{๐}^{}{}_{}{}^{๐}$$
(2.7)
The parameters $`\stackrel{\mathbf{~}}{๐}_{๐๐}`$ and $`\stackrel{\mathbf{~}}{๐}_{๐๐}`$ are defined by
$`\stackrel{\mathbf{~}}{๐}_{๐๐}`$ $`\mathbf{:=}`$ $`๐_{๐๐}\mathbf{+}\mathrm{๐}๐ฉ_{\mathbf{[}๐}^{}{}_{}{}^{๐}๐_{๐\mathbf{]}๐}\mathbf{+}๐ฉ_{๐}^{}{}_{}{}^{๐}๐ฉ_{๐}^{}{}_{}{}^{๐}๐_{๐๐}`$
$`\stackrel{\mathbf{~}}{๐}_{๐๐}`$ $`\mathbf{:=}`$ $`๐_{๐๐}\mathbf{}๐ฉ_{๐}^{}{}_{}{}^{๐}๐_{๐๐}`$ (2.8)
It is easy to see that in the dimensionally reduced theory, one can make use of the parameter components $`\stackrel{\mathbf{~}}{๐}_{๐๐}`$ and $`\stackrel{\mathbf{~}}{๐}_{๐๐}`$ to set $`๐ฉ_{๐๐๐}\mathbf{=}๐ฉ_{๐๐๐}\mathbf{=}\mathrm{๐}`$. Since we have not been able so far to cast the supersymmetry variations of these components into a completely $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ covariant form, this gauge would also be very convenient in the present setting. However, there appears to be an obstacle to this gauge choice if the full $`๐
\mathbf{=}\mathrm{๐๐}`$ coordinate dependence is retained.
To identify the proper $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ covariant bosonic fields, we must first explain how to rewrite the fermion fields. The $`๐
\mathbf{=}\mathrm{๐๐}`$ gravitino $`๐ฟ_๐จ\mathbf{}\mathbf{(}๐ฟ_{\underset{\mathbf{ยฏ}}{๐ถ}}\mathbf{,}๐ฟ_๐\mathbf{)}`$ has 32 spinor components, which split as $`\mathrm{๐}\mathbf{}\mathbf{(}\mathrm{๐}_๐\mathbf{}\mathrm{๐}_๐\mathbf{)}`$ under the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ subgroup of $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐๐}\mathbf{)}`$. Suppressing $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}`$ spinor indices, we then assign the resulting fields to the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ fields $`๐_๐^๐ฐ`$ and $`๐^{\dot{๐จ}}`$ via the following prescription
$$๐_๐^๐ฐ\mathbf{:=}\mathbf{\{}\begin{array}{cc}๐ซ^{\mathbf{}\mathrm{๐}\mathbf{/}\mathrm{๐}}๐_{๐}^{}{}_{}{}^{\underset{\mathbf{ยฏ}}{๐ถ}}\mathbf{\left(}๐ฟ_{\underset{\mathbf{ยฏ}}{๐ถ}๐ถ}\mathbf{+}๐ธ_{\underset{\mathbf{ยฏ}}{๐ถ}}๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{\right)}\hfill & \text{ if }๐ฐ\mathbf{=}๐ถ\hfill \\ ๐ซ^{\mathbf{}\mathrm{๐}\mathbf{/}\mathrm{๐}}๐_{๐}^{}{}_{}{}^{\underset{\mathbf{ยฏ}}{๐ถ}}\mathbf{\left(}๐ฟ_{\underset{\mathbf{ยฏ}}{๐ถ}\dot{๐ถ}}\mathbf{}๐ธ_{\underset{\mathbf{ยฏ}}{๐ถ}}๐ช_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{\right)}\hfill & \text{ if }๐ฐ\mathbf{=}\dot{๐ถ}\hfill \end{array}$$
(2.9)
and
$$๐^{\dot{๐จ}}\mathbf{:=}\mathbf{\{}\begin{array}{cc}๐ซ^{\mathbf{}\mathrm{๐}\mathbf{/}\mathrm{๐}}\mathbf{\left(}๐ช^๐๐ช^๐\mathbf{\right)}_{๐ถ๐ท}๐ฟ_{๐๐ท}\hfill & \text{ if }\dot{๐จ}\mathbf{=}\mathbf{(}๐๐ถ\mathbf{)}\hfill \\ \mathbf{}๐ซ^{\mathbf{}\mathrm{๐}\mathbf{/}\mathrm{๐}}\mathbf{\left(}๐ช^๐๐ช^๐\mathbf{\right)}_{\dot{๐ถ}\dot{๐ท}}๐ฟ_{๐\dot{๐ท}}\hfill & \text{ if }\dot{๐จ}\mathbf{=}\mathbf{(}\dot{๐ถ}๐\mathbf{)}\hfill \end{array}$$
(2.10)
where $`๐ฐ`$ and $`\dot{๐จ}`$ are $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ vector and (conjugate) spinor indices, respectively (see appendix B, and in particular (Appendix B:) for the relevant $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ decompositions).
The physical bosonic degrees of freedom, which correspond to the propagating 128 propagating scalar degrees of freedom of maximal $`๐
\mathbf{=}\mathrm{๐}`$ supergravity, are fused into an appropriate generalized vielbein. The relevant expressions are found by proceeding from the following $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ invariant ansatz for the supersymmetry variations of the vector fields in terms of the fermions (2.9) and (2.10) (in a suitable normalization):
$`๐น๐ฉ_๐^๐`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐ฐ๐ฑ}^๐\overline{\mathit{ฯต}}^๐ฐ๐_๐^๐ฑ\mathbf{+}๐_๐จ^๐๐ช_{๐จ\dot{๐ฉ}}^๐ฐ\overline{\mathit{ฯต}}^๐ฐ๐ธ_๐๐^{\dot{๐ฉ}}`$ (2.11)
$`๐น๐ฉ_{๐๐๐}`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐๐๐ฐ๐ฑ}\overline{\mathit{ฯต}}^๐ฐ๐_๐^๐ฑ\mathbf{+}๐_{๐๐๐จ}๐ช_{๐จ\dot{๐ฉ}}^๐ฐ\overline{\mathit{ฯต}}^๐ฐ๐ธ_๐๐^{\dot{๐ฉ}}`$ (2.12)
The explicit expressions in a special $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ gauge for the new bosonic quantities appearing on the r.h.s. of this equation can be found by comparing the above expressions with the ones obtained directly from $`๐
\mathbf{=}\mathrm{๐๐}`$ supergravity in the gauge (2.1). It is already known that
$$\mathbf{(}๐_{๐ฐ๐ฑ}^๐\mathbf{,}๐_๐จ^๐\mathbf{)}\mathbf{:=}\mathbf{\{}\begin{array}{cc}๐ซ^\mathbf{}\mathrm{๐}๐_๐{}_{}{}^{๐}๐ช_{๐ถ\dot{๐ท}}^{๐}\hfill & \text{if }\mathbf{[}๐ฐ๐ฑ\mathbf{]}\text{ or }๐จ\mathbf{=}\mathbf{(}๐ถ\dot{๐ท}\mathbf{)}\hfill \\ \mathrm{๐}\hfill & \text{otherwise}\hfill \end{array}$$
(2.13)
again using the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ decompositions of appendix B. By contrast, the objects $`\mathbf{(}๐_{๐๐๐ฐ๐ฑ}\mathbf{,}๐_{๐๐๐จ}\mathbf{)}`$ related to the gauge fields $`๐ฉ_{๐๐๐}`$ โ hence with antisymmetrized lower internal world indices โ have not yet appeared in previous work. Matching the r.h.s. of (2.12) with the variations obtained directly from the $`๐
\mathbf{=}\mathrm{๐๐}`$ supersymmetry variations of we find
$$๐_{๐๐๐}\mathbf{:=}\underset{๐๐๐}{\overset{\mathbf{}}{๐}}\mathbf{+}๐จ_{๐๐๐}๐_๐^๐$$
(2.14)
where
$$\underset{๐๐๐ฐ๐ฑ}{\overset{\mathbf{}}{๐}}\mathbf{:=}\mathbf{\{}\begin{array}{cc}๐ซ^\mathbf{}\mathrm{๐}๐_๐{}_{}{}^{๐}๐_{๐}^{}{}_{}{}^{๐}\mathbf{(}๐ช_{๐๐}\mathbf{)}_{๐ถ๐ท}^{}\hfill & \text{if }\mathbf{[}๐ฐ๐ฑ\mathbf{]}\mathbf{=}\mathbf{[}๐ถ๐ท\mathbf{]}\hfill \\ \mathbf{}๐ซ^\mathbf{}\mathrm{๐}๐_๐{}_{}{}^{๐}๐_{๐}^{}{}_{}{}^{๐}\mathbf{(}๐ช_{๐๐}\mathbf{)}_{\dot{๐ถ}\dot{๐ท}}^{}\hfill & \text{if }\mathbf{[}๐ฐ๐ฑ\mathbf{]}\mathbf{=}\mathbf{[}\dot{๐ถ}\dot{๐ท}\mathbf{]}\hfill \\ \mathrm{๐}\hfill & \text{if }\mathbf{[}๐ฐ๐ฑ\mathbf{]}\mathbf{=}\mathbf{[}๐ถ\dot{๐ท}\mathbf{]}\mathbf{}\mathbf{}\mathbf{[}\dot{๐ท}๐ถ\mathbf{]}\hfill \end{array}$$
(2.15)
$$\underset{๐๐๐จ}{\overset{\mathbf{}}{๐}}\mathbf{:=}\mathbf{\{}\begin{array}{cc}\mathrm{๐}๐ซ^\mathbf{}\mathrm{๐}๐\begin{array}{c}_{๐๐}๐_{๐๐}\hfill \\ \text{ }\text{ }\text{ }\hfill \end{array}\hfill & \text{if }๐จ\mathbf{=}\mathbf{(}๐๐\mathbf{)}\hfill \\ \mathrm{๐}\hfill & \text{if }๐จ\mathbf{=}\mathbf{(}๐ถ\dot{๐ท}\mathbf{)}\hfill \end{array}$$
(2.16)
Labeling the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ indices $`\mathbf{(}\mathbf{[}๐ฐ๐ฑ\mathbf{]}\mathbf{,}๐จ\mathbf{)}`$ collectively by $`๐\mathbf{,}๐\mathbf{,}\mathbf{}\mathbf{=}\mathrm{๐}\mathbf{,}\mathbf{}\mathbf{,}\mathrm{๐๐๐}`$ as in appendix A, the new objects $`๐_๐^๐`$ and $`๐_{๐๐๐}`$ together form a rectangular $`\mathbf{(}\mathrm{๐}\mathbf{+}\mathrm{๐๐}\mathbf{)}\mathbf{\times }\mathrm{๐๐๐}`$ matrix.
The new vielbeine are manifestly covariant w.r.t. local $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$, thereby enlarging the action of $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ of the original theory. While the supersymmetry variation of $`๐_๐^๐`$ was already given in , the variation of $`๐_{๐๐๐}`$ has so far not been determined. In appendix C we show that
$$๐น๐_{๐๐๐ฐ๐ฑ}\mathbf{=}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}๐^๐จ๐_{๐๐๐ฉ}๐น๐_{๐๐๐จ}\mathbf{=}\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}๐^๐ฉ๐_{๐๐๐ฐ๐ฑ}$$
(2.17)
with the local $`๐ต\mathbf{=}\mathrm{๐๐}`$ supersymmetry parameter
$$๐^๐จ\mathbf{:=}\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐จ\dot{๐จ}}^๐ฐ\overline{๐บ}^๐ฐ๐^{\dot{๐จ}}\mathbf{,}$$
(2.18)
In deriving this result, a compensating $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ rotation must be taken into account to restore the triangular gauge. It is an important consistency check that this compensating rotation comes out to be the same for $`๐_๐^๐`$ and $`๐_{๐๐๐}`$, and that the resulting transformation is exactly the same as for both fields, as required by consistency. We can therefore combine the supersymmetry variations of the generalized vielbeine with local $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ into the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ covariant form
$$๐น๐_๐^๐\mathbf{=}๐_{๐๐}^{}{}_{}{}^{๐}๐^๐๐_๐^๐๐น๐_{๐๐๐}\mathbf{=}๐_{๐๐}^{}{}_{}{}^{๐}๐^๐๐_{๐๐๐}$$
(2.19)
We also note that the components $`๐_{๐๐๐}`$ were not needed in because they cannot appear in the supersymmetry variations of the fermionic fields (the latter depend on the 3-form potential via the 4-index field strengths only).
All fields transform under general coordinate transformations in eleven dimensions. Splitting the $`๐
\mathbf{=}\mathrm{๐๐}`$ parameter as $`๐^๐ด\mathbf{=}\mathbf{(}๐^๐\mathbf{,}๐^๐\mathbf{)}`$, the transformations generated by $`๐^๐`$ take the standard form. For the remaining โinternalโ coordinate transformations with parameter $`๐^๐`$, we have
$`๐น๐ฉ_{๐}^{}{}_{}{}^{๐}`$ $`\mathbf{=}`$ $`๐_๐๐^๐\mathbf{+}๐^๐\mathbf{}_๐๐ฉ_{๐}^{}{}_{}{}^{๐}`$
$`๐น๐ฉ_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐^๐๐ฉ_{๐๐\mathbf{]}๐}\mathbf{+}๐^๐\mathbf{}_๐๐ฉ_{๐๐๐}`$ (2.20)
Furthermore,
$`๐น๐_๐^๐`$ $`\mathbf{=}`$ $`๐^๐\mathbf{}_๐๐_๐^๐\mathbf{}\mathbf{}_๐๐^๐๐_๐^๐\mathbf{}\mathbf{}_๐๐^๐๐_๐^๐`$
$`๐น๐_{๐๐๐}`$ $`\mathbf{=}`$ $`๐^๐\mathbf{}_๐๐_{๐๐๐}\mathbf{+}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐^๐๐_{๐\mathbf{]}๐๐}\mathbf{}\mathbf{}_๐๐^๐๐_{๐๐๐}`$ (2.21)
Whereas all the $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ fields considered previously were inert under tensor gauge transformations $`๐น๐จ_{๐ด๐ต๐ท}\mathbf{=}\mathrm{๐}\mathbf{}_{\mathbf{[}๐ด}๐_{๐ต๐ท\mathbf{]}}`$, the non-invariance of $`๐_{๐๐๐}`$ under such transformations, due to the appearance of the 3-index field $`๐จ_{๐๐๐}`$ in its definition, is a new feature. Specifically, under tensor gauge transformations with parameter $`๐_{๐๐}`$ we have
$`๐น๐ฉ_{๐}^{}{}_{}{}^{๐}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐น๐ฉ_{๐๐๐}`$ $`\mathbf{=}`$ $`๐_๐๐_{๐๐}\mathbf{}\mathrm{๐}๐ฉ_{๐}^{}{}_{}{}^{๐}\mathbf{}_{\mathbf{[}๐}๐_{๐\mathbf{]}๐}`$ (2.22)
and
$`๐น๐_๐^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐น๐_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathbf{}_๐๐_{๐๐}๐_๐^๐\mathbf{+}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐_{๐\mathbf{]}๐}๐_๐^๐`$ (2.23)
When combined with the previous coordinate transformations, these formulas are very suggestive of a unification of the internal coordinate transformations and tensor gauge transformations, based on combining the internal coordinate transformation parameters $`๐^๐`$ with the residual tensor gauge parameters $`๐_{๐๐}`$ into a single set $`\mathbf{(}๐^๐\mathbf{,}๐_{๐๐}\mathbf{)}`$ of 36 parameters. In the remaining sections we will show how these local symmetries are related to the the maximal nilpotent commuting subalgebra of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$.
## 3 Solution of algebraic constraints on the generalized vielbein
From the expressions (2.13)-(2.16) one can deduce a number of algebraic constraints on the generalized vielbein. They are
$`๐_๐จ^๐๐_๐จ^๐\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐ฐ๐ฑ}^๐๐_{๐ฐ๐ฑ}^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.1)
$`๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}\mathbf{\left(}๐_๐ฉ^๐๐_{๐ฐ๐ฑ}^๐\mathbf{}๐_๐ฉ^๐๐_{๐ฐ๐ฑ}^๐\mathbf{\right)}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}๐_๐จ^๐๐_๐ฉ^๐\mathbf{+}\mathrm{๐}๐_{๐ฒ\mathbf{[}๐ฐ}^๐๐_{๐ฑ\mathbf{]}๐ฒ}^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.2)
and
$`๐_{๐ฐ๐ฒ}^{\mathbf{(}๐}๐_{๐ฑ๐ฒ}^{๐\mathbf{)}}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐น_{๐ฐ๐ฑ}๐_{๐ฒ๐ณ}^๐๐_{๐ฒ๐ณ}^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐_{\mathbf{[}๐ฐ๐ฑ}^{\mathbf{(}๐}๐_{๐ฒ๐ณ\mathbf{]}}^{๐\mathbf{)}}\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐_๐จ^๐๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ๐ฒ๐ณ}๐_๐ฉ^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐ช_{\dot{๐จ}๐ฉ}^๐ฒ๐_๐ฉ^{\mathbf{(}๐}๐_{๐ฒ๐ณ}^{๐\mathbf{)}}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐ช_{\dot{๐จ}๐ฉ}^{๐ฐ๐ฒ๐ณ}๐_๐ฉ^{\mathbf{(}๐}๐_{๐ฒ๐ณ}^{๐\mathbf{)}}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.3)
For the new components, we find
$`๐_{๐๐๐จ}๐_{๐๐๐จ}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐๐๐ฐ๐ฑ}๐_{๐๐๐ฐ๐ฑ}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.4)
$`๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}\mathbf{\left(}๐_{๐๐๐ฉ}๐_{๐๐๐ฐ๐ฑ}\mathbf{}๐_{๐๐๐ฉ}๐_{๐๐๐ฐ๐ฑ}\mathbf{\right)}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}๐_{๐๐๐จ}๐_{๐๐๐ฉ}\mathbf{+}\mathrm{๐}๐_{๐๐๐ฒ\mathbf{[}๐ฐ}๐_{๐๐๐ฑ\mathbf{]}๐ฒ}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.5)
and
$`๐_{๐๐๐จ}๐_๐จ^๐\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐๐๐ฐ๐ฑ}๐_{๐ฐ๐ฑ}^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.6)
$`๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}\mathbf{\left(}๐_{๐๐๐ฉ}๐_{๐ฐ๐ฑ}^๐\mathbf{}๐_๐ฉ^๐๐_{๐๐๐ฐ๐ฑ}\mathbf{\right)}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}๐_{๐๐๐จ}๐_๐ฉ^๐\mathbf{+}\mathrm{๐}๐_{๐๐๐ฒ\mathbf{[}๐ฐ}๐_{๐ฑ\mathbf{]}๐ฒ}^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.7)
whereas no analog of (3.3) exists for the components $`๐_{๐๐๐}`$. All these relations are proved by decomposing the $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ into $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ and verifying the vanishing of their components case by case.
The above constraints can be elegantly rewritten in an $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ covariant form by means of the projectors onto the invariant subspaces of the tensor product of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ representations $`\mathrm{๐๐๐}\mathbf{}\mathrm{๐๐๐}`$ w.r.t. the decomposition $`\mathrm{๐๐๐}\mathbf{}\mathrm{๐๐๐}\mathbf{=}\mathrm{๐}\mathbf{}\mathrm{๐๐๐}\mathbf{}\mathrm{๐๐๐๐}\mathbf{}\mathrm{๐๐๐๐๐}\mathbf{}\mathrm{๐๐๐๐๐}`$. The relevant projectors are explicitly given in terms of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ structure constants by
$`\mathbf{(}๐_\mathrm{๐}\mathbf{)}_{๐๐}^{๐๐}`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐๐๐}}๐ผ_{๐๐}๐ผ^{๐๐}\mathbf{,}`$ (3.8)
$`\mathbf{(}๐_{\mathrm{๐๐๐}}\mathbf{)}_{๐๐}^{๐๐}`$ $`\mathbf{=}`$ $`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐^๐{}_{๐๐}{}^{}๐_{๐}^{}{}_{}{}^{๐๐}\mathbf{,}`$
$`\mathbf{(}๐_{\mathrm{๐๐๐๐}}\mathbf{)}_{๐๐}^{๐๐}`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}๐น_{\mathbf{(}๐}^๐๐น_{๐\mathbf{)}}^๐\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐ผ_{๐๐}๐ผ^{๐๐}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐^๐{}_{๐}{}^{}{}_{}{}^{\mathbf{(}๐}๐_{๐๐}^{}{}_{}{}^{๐\mathbf{)}}\mathbf{,}`$ (3.9)
It is straightforward to see that (3.1), (3.2) (3.4)-(3.7) are equivalent to
$`\mathbf{(}๐_๐\mathbf{)}_{๐๐}{}_{}{}^{๐๐}๐_{๐}^{๐}๐_๐^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`\mathbf{(}๐_๐\mathbf{)}_{๐๐}{}_{}{}^{๐๐}๐_{๐}^{๐}๐_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`\mathbf{(}๐_๐\mathbf{)}_{๐๐}{}_{}{}^{๐๐}๐_{๐๐๐}^{}๐_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathrm{๐}`$ (3.10)
for $`๐\mathbf{=}\mathrm{๐}`$ and $`\mathrm{๐๐๐}`$. It takes a little more work to verify that (3.3) can be expressed in the form<sup>3</sup><sup>3</sup>3A further relation given in
$$e_A^{(m}e_B^{n)}+\frac{1}{32}(\mathrm{\Gamma }^{IJ}\mathrm{\Gamma }^{KL})_{AB}e_{IJ}^me_{KL}^n\frac{1}{16}\mathrm{\Gamma }_{AC}^{IJ}e_C^{(m}e_D^{n)}\mathrm{\Gamma }_{DB}^{KL}=0$$
can be shown to be equivalent to (3.3) by means of a Fierz transformation and yields no new information. Therefore, there are no relations involving the projectors $`๐ซ_{27000}`$ and $`๐ซ_{30380}`$.
$$\mathbf{(}๐_{\mathrm{๐๐๐๐}}\mathbf{)}_{๐๐}{}_{}{}^{๐๐}๐_{๐}^{๐}๐_๐^๐\mathbf{=}\mathrm{๐}$$
(3.11)
Observe the invariance of the constraints (3) and (3.11) under the combined general coordinate and tensor gauge transformations (2) and (2): the transformed generalized vielbein still obeys all constraints.
We next demonstrate that the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ invariant algebraic relations on the generalized vielbein given above can be solved in terms of an $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ matrix $`๐ฅ`$. For the dimensionally reduced theory this is, of course, the expected result ), but with the important difference that the dependence on all eleven coordinates is here retained. Thus, $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ is already present in eleven dimensions, though not a symmetry of the theory. The existence of $`๐ฅ`$ also clarifies why we end up with the right number of physical degrees of freedom, a fact that cannot be directly ascertained by counting the constraints: being subject to local $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ transformations, the matrix $`๐ฅ`$ possesses just the $`\mathrm{๐๐๐}\mathbf{=}\mathrm{๐๐๐}\mathbf{}\mathrm{๐๐๐}`$ degrees of freedom of the coset $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}\mathbf{/}๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$. Thus, the counting works exactly as for the reduced theory.
To corroborate this claim, consider the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ Lie algebra valued matrices
$`\stackrel{\mathbf{~}}{๐ฌ}^๐`$ $`\mathbf{:=}`$ $`๐_๐^๐๐ง^๐\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐ฐ๐ฑ}^๐๐ฟ^{๐ฐ๐ฑ}\mathbf{+}๐_๐จ^๐๐^๐จ`$
$`\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}`$ $`\mathbf{:=}`$ $`๐_{๐๐๐}๐ง^๐\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐๐๐ฐ๐ฑ}๐ฟ^{๐ฐ๐ฑ}\mathbf{+}๐_{๐๐๐จ}๐^๐จ`$ (3.12)
From the relations presented in the foregoing section we infer that these matrices commute (cf. (3.2), (3.5), (3.7))
$$\mathbf{[}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{,}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{]}\mathbf{=}\mathbf{[}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{,}\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{]}\mathbf{=}\mathbf{[}\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{,}\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{]}\mathbf{=}\mathrm{๐}$$
(3.13)
and are nilpotent (cf. (3.1), (3.4), (3.6))
$$\mathrm{๐๐ซ}\mathbf{\left(}\stackrel{\mathbf{~}}{๐ฌ}^๐\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{\right)}\mathbf{=}\mathrm{๐๐ซ}\mathbf{\left(}\stackrel{\mathbf{~}}{๐ฌ}^๐\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{\right)}\mathbf{=}\mathrm{๐๐ซ}\mathbf{\left(}\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{\right)}\mathbf{=}\mathrm{๐}$$
(3.14)
(ensuring that any linear combination of these matrices has norm zero). Since they are linearly independent (as is most easily checked by setting $`๐_{๐}^{}{}_{}{}^{๐}\mathbf{=}๐น_๐^๐`$ in the original definition), they form a 36 dimensional abelian nilpotent subalgebra of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$. There is only one such algebra, which is unique up to conjugation . Consequently, there must exist an $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ matrix $`๐ฅ`$ such that
$$\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{=}๐ฅ^\mathbf{}\mathrm{๐}๐^๐๐ฅ\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{=}๐ฅ^\mathbf{}\mathrm{๐}๐_{๐๐}๐ฅ$$
(3.15)
where $`๐^๐`$ and $`๐_{๐๐}`$ are the 8+28 nilpotent generators of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ introduced in the appendix. The assignment of the 8+28 vielbein components to these generators here is uniquely determined by $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ covariance; its correctness will be confirmed below when we analyze (3.3). Thus,
$`๐_๐^๐`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐๐}}\mathrm{๐๐ซ}\mathbf{(}๐^๐๐ฅ๐ง_๐๐ฅ^\mathbf{}\mathrm{๐}\mathbf{)}`$
$`๐_{๐๐๐}`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐๐}}\mathrm{๐๐ซ}\mathbf{(}๐_{๐๐}๐ฅ๐ง_๐๐ฅ^\mathbf{}\mathrm{๐}\mathbf{)}`$ (3.16)
By use of the relation
$$\frac{\mathrm{๐}}{\mathrm{๐๐}}\mathrm{๐๐ซ}\mathbf{\left(}๐ง^๐๐ฅ๐ง_๐๐ฅ^\mathbf{}\mathrm{๐}\mathbf{\right)}\mathbf{=}๐ฅ_{}^{๐}{}_{๐}{}^{}$$
(3.17)
for the adjoint representation we can write
$$๐_๐^๐\mathbf{=}๐ฅ_{}^{๐}{}_{๐}{}^{}๐_{๐๐๐}\mathbf{=}๐ฅ_{๐๐๐}$$
(3.18)
whence the generalized vielbein $`\mathbf{(}๐_๐^๐\mathbf{,}๐_{๐๐๐}\mathbf{)}`$ is actually a rectangular submatrix of $`๐ฅ`$. Let us emphasize once more that $`๐ฅ`$ still depends on eleven coordinates.
At this point, a remark concerning our use of indices is in order. In the appendix, we use โflatโ indices $`๐\mathbf{,}๐\mathbf{,}๐\mathbf{=}\mathrm{๐}\mathbf{,}\mathbf{}\mathbf{,}\mathrm{๐}`$ to label the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ generators in the $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ decomposition. On the other hand, the index $`๐`$ appearing on the l.h.s. of (2.13) should be viewed as โcurvedโ in the sense that it is acted on by internal coordinate transformations. Of course, there is no need for such a distinction in a flat background characterized by $`๐_๐{}_{}{}^{๐}\mathbf{=}๐น_๐^๐`$ and $`๐ฅ\mathbf{=}\mathrm{๐}`$, whereas the two kinds of indices no longer coincide for curved backgrounds characterized by non-trivial $`๐_๐^๐`$ and $`๐จ_{๐๐๐}`$. The nilpotent generators in (3.12) thus represent โcurvedโ analogs of the โflatโ generators $`๐^๐`$ and $`๐_{๐๐}`$, and the above relations tell us is that the transition from flat to curved configurations is entirely accounted for by the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ matrix $`๐ฅ`$. This illustrates the enlargement of the geometry in comparison with the conventional description $`๐
\mathbf{=}\mathrm{๐๐}`$ supergravity, where the achtbein can only be deformed with a $`๐ฎ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ matrix.
To confirm the consistency of the above solution let us analyze the third set of constraints (3.3), which we have not yet discussed. Inspection reveals that the desired relation is equivalent to
$$๐_{\mathrm{๐๐๐๐}}\mathbf{\left(}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{\right)}\mathbf{=}\mathrm{๐}$$
(3.19)
Making use of the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ invariance of $`๐_{\mathrm{๐๐๐๐}}`$, we can replace curved by flat indices in this relation. This yields
$$\mathbf{(}๐_{\mathrm{๐๐๐๐}}\mathbf{)}_{๐๐}^{}{}_{}{}^{๐๐
}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^{\mathbf{(}๐}๐น_๐^{๐
\mathbf{)}}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐ผ_{๐๐}๐ผ^{๐๐
}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐^๐{}_{๐}{}^{}{}_{}{}^{\mathbf{(}๐}๐_{๐๐}^{}{}_{}{}^{๐
\mathbf{)}}\mathbf{=}\mathrm{๐}$$
(3.20)
By nilpotency, we have $`๐ผ^{๐๐
}\mathbf{=}\mathrm{๐}`$, and contracting the remaining terms with $`๐ง^๐`$ we see that (2.13) does satisfy (3.3), provided the following relation holds for all $`๐`$
$$\mathbf{[}\mathbf{[}๐ง_๐\mathbf{,}๐^๐\mathbf{]}\mathbf{,}๐^๐
\mathbf{]}\mathbf{=}\mathbf{}\mathrm{๐}๐น_๐^{\mathbf{(}๐}๐^{๐
\mathbf{)}}$$
(3.21)
Since the algebra preserves the grading, the relation is trivially satisfied for all generators except $`๐ง_๐\mathbf{=}๐_๐`$, which must be checked separately. A quick calculation, using the commutation relations listed in appendix B, shows that the required relation is indeed satisfied. Let us emphasize once more that there is no analog of (3.20) for nilpotent elements conjugate to the $`๐_{๐๐}`$.
Having established the consistency of (3) it remains to investigate its uniqueness. It is easy to see that $`๐ฅ`$ is, in fact, not unique because (3) remains unchanged under the transformation
$$๐ฅ\mathbf{}๐\mathbf{}๐ฅ\mathrm{๐๐จ๐ซ}๐\mathbf{}๐$$
(3.22)
where $`๐`$ is the Borel subgroup of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ generated by $`๐^๐`$ and $`๐_{๐๐}`$. This remaining non-uniqueness can be fixed by invoking the differential relations
$`๐_๐จ^๐๐ท_๐^๐จ`$ $`\mathbf{=}`$ $`๐บ_{๐}^{}{}_{}{}^{๐๐}๐_{๐๐}^{}{}_{}{}^{๐}`$ (3.23)
$`๐_{๐๐๐จ}๐ท_๐^๐จ`$ $`\mathbf{=}`$ $`๐บ_{๐}^{}{}_{}{}^{๐๐}๐_{๐๐๐๐}`$ (3.24)
where $`๐_{๐๐}^{}{}_{}{}^{๐}`$ was already defined in (2.7), and
$$๐_{๐๐๐๐}\mathbf{:=}๐_๐๐ฉ_{๐๐๐}\mathbf{}๐_๐๐ฉ_{๐๐๐}\mathbf{+}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐ฉ_{\mathbf{[}๐}^{}{}_{}{}^{๐}๐ฉ_{๐\mathbf{]}๐\mathbf{]}๐}\mathbf{+}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐ฉ_{๐\mathbf{]}๐๐}$$
(3.25)
(3.23) thus relates the field strengths to (part of) the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ connection $`๐ท_๐^๐จ`$, whose explicit expression in terms of the spin connection and the 4-index field strength is given in formula (22) of . In the reduction to three dimensions these differential constraints become the linear duality relations that allow us to trade the 36 vector fields for their dual scalars.
## 4 Outlook
The matrix $`๐ฅ`$ plays a role similar to the achtbein $`๐_{๐}^{}{}_{}{}^{๐}`$, but also incorporates the tensor degrees of freedom from $`๐จ_{๐๐๐}`$, as well as the vector fields $`๐ฉ_{๐}^{}{}_{}{}^{๐}`$ and $`๐ฉ_{๐๐๐}`$. We would now like to argue that $`๐ฅ\mathbf{}๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}\mathbf{/}๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ really is the appropriate vielbein encompassing the propagating degrees of freedom of $`๐
\mathbf{=}\mathrm{๐๐}`$ supergravity, in the same way as the ordinary vielbein, viewed as an element of $`๐ฎ๐ณ\mathbf{(}๐
\mathbf{,}\mathbf{)}\mathbf{/}๐บ๐ถ\mathbf{(}๐
\mathbf{)}`$, describes the graviton degrees of freedom (with a Euclidean signature). To this aim, let us first recall that the internal part of the (inverse densitized) metric is recovered from the generalized vielbein via the $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ invariant formula
$$๐ซ^\mathbf{}\mathrm{๐}๐^{๐๐}\mathbf{=}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐_๐จ^๐๐_๐จ^๐\mathbf{=}\frac{\mathrm{๐}}{\mathrm{๐๐๐}}\mathbf{\left(}๐_๐จ^๐๐_๐จ^๐\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{๐ฐ๐ฑ}^๐๐_{๐ฐ๐ฑ}^๐\mathbf{\right)}$$
(4.1)
where the constraint (3.1) was used. The summation on the r.h.s. breaks $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ to $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ because of the plus sign in front of the second term; with a minus sign, the expression would vanish by the constraints.
Just as for the standard vielbein, and having already introduced this terminology in the foregoing section, we now interpret the indices $`๐\mathbf{,}๐\mathbf{,}\mathbf{}`$ and $`๐\mathbf{,}๐\mathbf{,}\mathbf{}`$ appearing in (3.18) as โcurvedโ and โflatโ, respectively. This then immediately suggest the the following generalization of (4.1)
$$๐^{๐๐}\mathbf{:=}\frac{\mathrm{๐}}{\mathrm{๐๐๐}}๐ฅ_{}^{๐}{}_{๐}{}^{}๐ฅ_{}^{๐}{}_{๐}{}^{}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐๐}}\mathbf{\left(}๐ฅ_{}^{๐}{}_{๐จ}{}^{}๐ฅ_{}^{๐}{}_{๐จ}{}^{}\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐ฅ_{}^{๐}{}_{๐ฐ๐ฑ}{}^{}๐ฅ_{}^{๐}{}_{๐ฐ๐ฑ}{}^{}\mathbf{\right)}$$
(4.2)
By construction, this metric is invariant under local $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$, which acts as
$$๐ฅ_{}^{๐}{}_{๐}{}^{}\mathbf{}๐ฅ_{}^{๐}{}_{๐}{}^{}๐บ_{}^{๐}{}_{๐}{}^{}$$
(4.3)
on the 248-bein with $`๐บ`$ in the $`\mathrm{๐๐๐}\mathbf{}\mathrm{๐๐๐}`$ representation of $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ (and depending on all eleven coordinates). As we showed before, a โbosonizedโ version of local supersymmetry formally extends this to an action of a full local $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ acting from the right on $`๐ฅ_{}^{๐}{}_{๐}{}^{}`$ (which, however, does not leave $`๐^{๐๐}`$ inert any more).
Similarly, we now have a combined action of the internal coordinate and tensor gauge transformations on the 248-bein $`๐ฅ_{}^{๐}{}_{๐}{}^{}`$, which is analogous to the action of general coordinate transformations on the standard vielbein. Namely, from the previous formulas (2) and (2) we can directly read off their action on the $`\mathrm{๐๐}\mathbf{\times }\mathrm{๐๐๐}`$ submatrix of $`๐ฅ_{}^{๐}{}_{๐}{}^{}`$:
$`๐น๐ฅ_๐^๐`$ $`\mathbf{=}`$ $`๐^๐\mathbf{}_๐๐ฅ_๐^๐\mathbf{}\mathbf{}_๐๐^๐๐ฅ_๐^๐\mathbf{}\mathbf{}_๐๐^๐๐ฅ_๐^๐`$
$`๐น๐ฅ_{๐๐๐}`$ $`\mathbf{=}`$ $`๐^๐\mathbf{}_๐๐ฅ_{๐๐๐}\mathbf{}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐^๐๐ฅ_{๐\mathbf{]}๐๐}\mathbf{}\mathbf{}_๐๐^๐๐ฅ_{๐๐๐}`$ (4.4)
and
$`๐น๐ฅ_๐^๐`$ $`\mathbf{=}`$ $`\mathrm{๐}`$
$`๐น๐ฅ_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathbf{}_๐๐_{๐๐}๐ฅ_๐^๐\mathbf{+}\mathrm{๐}\mathbf{}_{\mathbf{[}๐}๐_{๐\mathbf{]}๐}๐ฅ_๐^๐`$ (4.5)
This action can be extended to the full 248-bein via
$`๐น๐ฅ`$ $`\mathbf{=}`$ $`๐^๐\mathbf{}_๐๐ฅ\mathbf{+}\mathbf{}_๐๐^๐\mathbf{\left(}๐ฌ_{๐}^{}{}_{}{}^{๐}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐ต\mathbf{\right)}๐ฅ`$
$`๐น๐ฅ`$ $`\mathbf{=}`$ $`\mathbf{}_๐๐_{๐๐}๐ฌ^{๐๐๐}๐ฅ`$ (4.6)
by means of the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ Lie algebra matrices given in appendix A (although this is usually not done, the general coordinate transformations on the standard vielbein can be cast into an analogous form by use of $`๐ฎ๐ณ\mathbf{(}๐
\mathbf{,}\mathbf{)}`$ matrices). It is important here that this action manifestly preserves $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$: the transformed vielbein is still an element of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$. It is also straightforward to exponentiate the infinitesimal action to a full โdiffeomorphismโ generated by the pair $`\mathbf{(}๐^๐\mathbf{,}๐_{๐๐}\mathbf{)}`$. Intriguingly, the missing โtransport termโ with $`๐_{๐๐}`$ suggests a hidden dependence on 28 further coordinates $`๐_{๐๐}`$, but it remains to be seen whether such an extension exists. Finally, it is clear that the rigid $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ invariance of the dimensionally reduced theory emerges from the above local symmetries in the same way as rigid $`๐ฎ๐ณ\mathbf{(}๐
\mathbf{,}\mathbf{)}`$ symmetry emerges in the torus reduction as a remnant of general coordinate invariance.
## Appendix Appendix A: The $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ decomposition of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$.
In the standard $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ decomposition $`\mathrm{๐๐๐}\mathbf{}\mathrm{๐๐๐}\mathbf{}\mathrm{๐๐๐}`$, the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ Lie algebra generators are $`๐ง^๐\mathbf{=}\mathbf{(}๐ฟ^{๐ฐ๐ฑ}\mathbf{,}๐^๐จ\mathbf{)}`$, with $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ vector and spinor indices $`๐ฐ\mathbf{,}๐ฑ\mathbf{=}\mathrm{๐}\mathbf{,}\mathbf{}\mathbf{,}\mathrm{๐๐}`$ and $`๐จ\mathbf{=}\mathrm{๐}\mathbf{,}\mathbf{}\mathbf{,}\mathrm{๐๐๐}`$, respectively. They obey
$`\mathbf{[}๐ฟ^{๐ฐ๐ฑ}\mathbf{,}๐ฟ^{๐ฒ๐ณ}\mathbf{]}`$ $`\mathbf{=}`$ $`\mathrm{๐}๐น\begin{array}{c}\text{ }\text{ }\text{ }\hfill \\ ^{๐ฐ\mathbf{[}๐ฒ}๐ฟ^{๐ณ\mathbf{]}๐ฑ}\hfill \end{array}`$ (A.3)
$`\mathbf{[}๐ฟ^{๐ฐ๐ฑ}\mathbf{,}๐^๐จ\mathbf{]}`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}๐ฟ^{๐ฐ๐ฑ}`$
$`\mathbf{[}๐^๐จ\mathbf{,}๐^๐ฉ\mathbf{]}`$ $`\mathbf{=}`$ $`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}๐^๐ฉ`$ (A.4)
The totally antisymmetric $`๐ฌ_\mathrm{๐}`$ structure constants $`๐^{๐๐๐}`$ therefore possess the non-vanishing components
$$๐^{๐ฐ๐ฑ\mathbf{,}๐ฒ๐ณ\mathbf{,}๐ด๐ต}\mathbf{=}\mathbf{}\mathrm{๐}๐น\begin{array}{c}\text{ }\text{ }\text{ }\hfill \\ ^{๐ฐ\mathbf{[}๐ฒ}_{}๐น_{๐ด๐ต}^{๐ณ\mathbf{]}๐ฑ}\hfill \end{array}\mathbf{,}๐^{๐ฐ๐ฑ\mathbf{,}๐จ\mathbf{,}๐ฉ}\mathbf{=}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐จ๐ฉ}^{๐ฐ๐ฑ}$$
(A.5)
The Cartan-Killing form is given by
$$๐ผ^{๐๐}\mathbf{=}\frac{\mathrm{๐}}{\mathrm{๐๐}}\mathrm{๐๐ซ}๐ง^๐๐ง^๐\mathbf{=}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}๐^๐{}_{๐๐}{}^{}๐_{}^{๐๐๐}$$
(A.6)
with components $`๐ผ^{๐จ๐ฉ}\mathbf{=}๐น^{๐จ๐ฉ}`$ and $`๐ผ^{๐ฐ๐ฑ๐ฒ๐ณ}\mathbf{=}\mathbf{}\mathrm{๐}๐น_{๐ฒ๐ณ}^{๐ฐ๐ฑ}`$. When summing over antisymmetrized index pairs $`\mathbf{[}๐ฐ๐ฑ\mathbf{]}`$, an extra factor of $`\frac{\mathrm{๐}}{\mathrm{๐}}`$ is always understood.
## Appendix Appendix B: The $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ decomposition of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$.
To recover the $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ basis of , we will further decompose the above representations into representations of the subgroup $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}\mathbf{}\mathbf{\left(}๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}\mathbf{\times }๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}\mathbf{\right)}_{\mathrm{๐๐ข๐๐ }}\mathbf{}๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$. The indices corresponding to the $`\mathrm{๐}_๐\mathbf{,}\mathrm{๐}_๐`$ and $`\mathrm{๐}_๐`$ representations of $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$, respectively, will be denoted by $`๐`$, $`๐ถ`$ and $`\dot{๐ถ}`$. After a triality rotation the $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ vector and spinor representations decompose as
$`\mathrm{๐๐}`$ $`\mathbf{}`$ $`\mathrm{๐}_๐ฌ\mathbf{}\mathrm{๐}_๐`$
$`\mathrm{๐๐๐}_๐ฌ`$ $`\mathbf{}`$ $`\mathbf{(}\mathrm{๐}_๐ฌ\mathbf{}\mathrm{๐}_๐\mathbf{)}\mathbf{}\mathbf{(}\mathrm{๐}_๐ฏ\mathbf{}\mathrm{๐}_๐ฏ\mathbf{)}\mathbf{=}\mathrm{๐}_๐ฏ\mathbf{}\mathrm{๐๐}_๐ฏ\mathbf{}\mathrm{๐}\mathbf{}\mathrm{๐๐}\mathbf{}\mathrm{๐๐}_๐ฏ`$
$`\mathrm{๐๐๐}_๐`$ $`\mathbf{}`$ $`\mathbf{(}\mathrm{๐}_๐ฏ\mathbf{}\mathrm{๐}_๐ฌ\mathbf{)}\mathbf{}\mathbf{(}\mathrm{๐}_๐\mathbf{}\mathrm{๐}_๐ฏ\mathbf{)}\mathbf{=}\mathrm{๐}_๐ฌ\mathbf{}\mathrm{๐๐}_๐ฌ\mathbf{}\mathrm{๐}_๐\mathbf{}\mathrm{๐๐}_๐\mathbf{,}`$ (B.1)
respectively. We thus have $`๐ฐ\mathbf{=}\mathbf{(}๐ถ\mathbf{,}\dot{๐ถ}\mathbf{)}`$ and $`๐จ\mathbf{=}\mathbf{(}๐ถ\dot{๐ท}\mathbf{,}๐๐\mathbf{)}`$, and the $`๐ฌ_\mathrm{๐}`$ generators decompose as
$$๐ฟ^{\mathbf{[}๐ฐ๐ฑ\mathbf{]}}\mathbf{}\mathbf{(}๐ฟ^{\mathbf{[}๐ถ๐ท\mathbf{]}}\mathbf{,}๐ฟ^{\mathbf{[}\dot{๐ถ}\dot{๐ท}\mathbf{]}}\mathbf{,}๐ฟ^{๐ถ\dot{๐ท}}\mathbf{)}\mathbf{,}๐^๐จ\mathbf{}\mathbf{(}๐^{๐ถ\dot{๐ถ}}\mathbf{,}๐^{๐๐}\mathbf{)}\mathbf{.}$$
(B.2)
Next we regroup these generators as follows. The 63 generators
$$๐ฌ_๐{}_{}{}^{๐}\mathbf{:=}\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{(}๐ช_{๐ถ๐ท}^{๐๐}๐ฟ^{\mathbf{[}๐ถ๐ท\mathbf{]}}\mathbf{+}๐ช_{\dot{๐ถ}\dot{๐ท}}^{๐๐}๐ฟ^{\mathbf{[}\dot{๐ถ}\dot{๐ท}\mathbf{]}}\mathbf{)}\mathbf{+}๐^{\mathbf{(}๐๐\mathbf{)}}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น^{๐๐}๐^{๐๐}\mathbf{,}$$
for $`\mathrm{๐}\mathbf{}๐\mathbf{,}๐\mathbf{}\mathrm{๐}`$ span an $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ subalgebra of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$. The generator
$$๐ต\mathbf{:=}๐^{๐๐}$$
extends this subalgebra to $`๐ฎ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$. The remainder of the $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ Lie algebra then decomposes into the following representations of $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$:
$`๐^๐`$ $`\mathbf{:=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐ถ\dot{๐ถ}}^๐\mathbf{(}๐ฟ^{๐ถ\dot{๐ถ}}\mathbf{+}๐^{๐ถ\dot{๐ถ}}\mathbf{)}\mathbf{,}`$
$`๐_{๐๐}`$ $`\mathbf{:=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{\left(}๐ช_{๐ถ๐ท}^{๐๐}๐ฟ^{\mathbf{[}๐ถ๐ท\mathbf{]}}\mathbf{}๐ช_{\dot{๐ถ}\dot{๐ท}}^{๐๐}๐ฟ^{\mathbf{[}\dot{๐ถ}\dot{๐ท}\mathbf{]}}\mathbf{\right)}\mathbf{+}๐^{\mathbf{[}๐๐\mathbf{]}}\mathbf{,}`$
$`๐ฌ^{๐๐๐}`$ $`\mathbf{:=}`$ $`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐ถ\dot{๐ถ}}^{๐๐๐}\mathbf{(}๐ฟ^{๐ถ\dot{๐ถ}}\mathbf{}๐^{๐ถ\dot{๐ถ}}\mathbf{)}`$ (B.3)
and
$`๐_๐`$ $`\mathbf{:=}`$ $`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐ถ\dot{๐ถ}}^๐\mathbf{(}๐ฟ^{๐ถ\dot{๐ถ}}\mathbf{}๐^{๐ถ\dot{๐ถ}}\mathbf{)}\mathbf{,}`$
$`๐^{๐๐}`$ $`\mathbf{:=}`$ $`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{(}๐ช_{๐ถ๐ท}^{๐๐}๐ฟ^{\mathbf{[}๐ถ๐ท\mathbf{]}}\mathbf{}๐ช_{\dot{๐ถ}\dot{๐ท}}^{๐๐}๐ฟ^{\mathbf{[}\dot{๐ถ}\dot{๐ท}\mathbf{]}}\mathbf{)}\mathbf{+}๐^{\mathbf{[}๐๐\mathbf{]}}\mathbf{,}`$
$`๐ฌ_{๐๐๐}`$ $`\mathbf{:=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}๐ช_{๐ถ\dot{๐ถ}}^{๐๐๐}\mathbf{(}๐ฟ^{๐ถ\dot{๐ถ}}\mathbf{+}๐^{๐ถ\dot{๐ถ}}\mathbf{)}\mathbf{,}`$ (B.4)
The Cartan subalgebra is spanned by the diagonal elements $`๐ฌ_\mathrm{๐}{}_{}{}^{\mathrm{๐}}\mathbf{,}\mathbf{}\mathbf{,}๐ฌ_\mathrm{๐}^\mathrm{๐}`$ and $`๐ต`$, or, equivalently, by $`๐^{\mathrm{๐๐}}\mathbf{,}\mathbf{}\mathbf{,}๐^{\mathrm{๐๐}}`$. Obviously, the elements $`๐ฌ_๐^๐`$ for $`๐\mathbf{<}๐`$ (or $`๐\mathbf{>}๐`$) together with the elements (B.3) (or (B.4)) for $`๐\mathbf{<}๐\mathbf{<}๐`$ generate the Borel subalgebra of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ associated with the positive (negative) roots of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$. Furthermore, these generators are graded w.r.t. the number of times the root $`๐ถ_\mathrm{๐}`$ (corresponding to the element $`๐ต`$ in the Cartan subalgebra) appears, such that for any basis generator $`๐ฟ`$ we have $`\mathbf{[}๐ต\mathbf{,}๐ฟ\mathbf{]}\mathbf{=}\mathrm{๐๐๐ }\mathbf{(}๐ฟ\mathbf{)}\mathbf{}๐ฟ`$. The degree can be read off from
$$\begin{array}{cccccc}\hfill \mathbf{[}๐ต\mathbf{,}๐^๐\mathbf{]}& \mathbf{=}& \mathrm{๐}๐^๐\hfill & \hfill \mathbf{[}๐ต\mathbf{,}๐_๐\mathbf{]}& \mathbf{=}& \mathbf{}\mathrm{๐}๐_๐\hfill \\ \hfill \mathbf{[}๐ต\mathbf{,}๐_{๐๐}\mathbf{]}& \mathbf{=}& \mathrm{๐}๐_{๐๐}\hfill & \hfill \mathbf{[}๐ต\mathbf{,}๐^{๐๐}\mathbf{]}& \mathbf{=}& \mathbf{}\mathrm{๐}๐^{๐๐}\hfill \\ \hfill \mathbf{[}๐ต\mathbf{,}๐ฌ^{๐๐๐}\mathbf{]}& \mathbf{=}& ๐ฌ^{๐๐๐}\hfill & \hfill \mathbf{[}๐ต\mathbf{,}๐ฌ_{๐๐๐}\mathbf{]}& \mathbf{=}& \mathbf{}๐ฌ_{๐๐๐}\hfill \\ \hfill \mathbf{[}๐ต\mathbf{,}๐ฌ_๐{}_{}{}^{๐}\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill & & & \end{array}$$
The remaining commutation relations are given by
$$\begin{array}{cccccc}\hfill \mathbf{[}๐^๐\mathbf{,}๐^๐\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill & \hfill \mathbf{[}๐_๐\mathbf{,}๐_๐\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill \\ \hfill \mathbf{[}๐_๐\mathbf{,}๐^๐\mathbf{]}& \mathbf{=}& ๐ฌ_๐{}_{}{}^{๐}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐ต\hfill & & & \end{array}$$
$$\begin{array}{cccccc}\hfill \mathbf{[}๐_{๐๐}\mathbf{,}๐^๐\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill & \hfill \mathbf{[}๐_{๐๐}\mathbf{,}๐_๐\mathbf{]}& \mathbf{=}& \mathbf{}๐ฌ_{๐๐๐}\hfill \\ \hfill \mathbf{[}๐_{๐๐}\mathbf{,}๐_{๐๐
}\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill & \hfill \mathbf{[}๐_{๐๐}\mathbf{,}๐^{๐๐
}\mathbf{]}& \mathbf{=}& \mathrm{๐}๐น_{\mathbf{[}๐}^{\mathbf{[}๐}๐ฌ_{๐\mathbf{]}}^{}{}_{}{}^{๐
\mathbf{]}}\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_{๐๐}^{๐๐
}๐ต\hfill \\ \hfill \mathbf{[}๐^{๐๐}\mathbf{,}๐^๐\mathbf{]}& \mathbf{=}& \mathbf{}๐ฌ^{๐๐๐}\hfill & \hfill \mathbf{[}๐^{๐๐}\mathbf{,}๐_๐\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill \end{array}$$
$$\begin{array}{cccccc}\hfill \mathbf{[}๐ฌ^{๐๐๐}\mathbf{,}๐^๐
\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill & \hfill \mathbf{[}๐ฌ_{๐๐๐}\mathbf{,}๐^๐
\mathbf{]}& \mathbf{=}& \mathrm{๐}๐น_{\mathbf{[}๐}^๐
๐_{๐๐\mathbf{]}}^{}\hfill \\ \hfill \mathbf{[}๐ฌ^{๐๐๐}\mathbf{,}๐_{๐
๐}\mathbf{]}& \mathbf{=}& \mathbf{}\mathrm{๐}๐น_{๐
๐}^{\mathbf{[}๐๐}๐_{}^{๐\mathbf{]}}\hfill & \hfill \mathbf{[}๐ฌ_{๐๐๐}\mathbf{,}๐_{๐
๐}\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill \\ \hfill \mathbf{[}๐ฌ^{๐๐๐}\mathbf{,}๐ฌ^{๐
๐๐}\mathbf{]}& \mathbf{=}& \mathbf{}\frac{\mathrm{๐}}{\mathrm{๐๐}}\mathit{ฯต}^{๐๐๐๐
๐๐๐๐}๐_{๐๐}\hfill & \hfill \mathbf{[}๐ฌ_{๐๐๐}\mathbf{,}๐ฌ_{๐
๐๐}\mathbf{]}& \mathbf{=}& \frac{\mathrm{๐}}{\mathrm{๐๐}}\mathit{ฯต}_{๐๐๐๐
๐๐๐๐}๐^{๐๐}\hfill \\ \hfill \mathbf{[}๐ฌ^{๐๐๐}\mathbf{,}๐_๐
\mathbf{]}& \mathbf{=}& \mathrm{๐}๐น_๐
^{\mathbf{[}๐}๐_{}^{๐๐\mathbf{]}}\hfill & \hfill \mathbf{[}๐ฌ_{๐๐๐}\mathbf{,}๐_๐
\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill \\ \hfill \mathbf{[}๐ฌ^{๐๐๐}\mathbf{,}๐^{๐
๐}\mathbf{]}& \mathbf{=}& \mathrm{๐}\hfill & \hfill \mathbf{[}๐ฌ_{๐๐๐}\mathbf{,}๐^{๐
๐}\mathbf{]}& \mathbf{=}& \mathrm{๐}๐น_{\mathbf{[}๐๐}^{๐
๐}๐_{๐\mathbf{]}}^{}\hfill \\ \hfill \mathbf{[}๐ฌ^{๐๐๐}\mathbf{,}๐ฌ_{๐
๐๐}\mathbf{]}& \mathbf{=}& \mathbf{}\mathrm{๐๐}๐น_{\mathbf{[}๐
๐}^{\mathbf{[}๐๐}๐ฌ_{๐\mathbf{]}}^{}{}_{}{}^{๐\mathbf{]}}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_{๐
๐๐}^{๐๐๐}๐ต\hfill & & & \end{array}$$
$$\begin{array}{cccccc}\hfill \mathbf{[}๐ฌ_๐{}_{}{}^{๐}\mathbf{,}๐^๐\mathbf{]}& \mathbf{=}& \mathbf{}๐น_๐^๐๐^๐\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐^๐\hfill & \hfill \mathbf{[}๐ฌ_๐{}_{}{}^{๐}\mathbf{,}๐_๐\mathbf{]}& \mathbf{=}& ๐น_๐^๐๐_๐\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐_๐\hfill \\ \hfill \mathbf{[}๐ฌ_๐{}_{}{}^{๐}\mathbf{,}๐_{๐๐
}\mathbf{]}& \mathbf{=}& \mathbf{}\mathrm{๐}๐น_{\mathbf{[}๐}^๐๐_{๐
\mathbf{]}๐}^{}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐_{๐๐
}\hfill & \hfill \mathbf{[}๐ฌ_๐{}_{}{}^{๐}\mathbf{,}๐^{๐๐
}\mathbf{]}& \mathbf{=}& \mathrm{๐}๐น_๐^{\mathbf{[}๐}๐_{}^{๐
\mathbf{]}๐}\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐^{๐๐
}\hfill \\ \hfill \mathbf{[}๐ฌ_๐{}_{}{}^{๐}\mathbf{,}๐ฌ^{๐๐
๐}\mathbf{]}& \mathbf{=}& \mathbf{}\mathrm{๐}๐น_๐^{\mathbf{[}๐}๐ฌ_{}^{๐
๐\mathbf{]}๐}\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐ฌ^{๐๐
๐}\hfill & \hfill \mathbf{[}๐ฌ_๐{}_{}{}^{๐}\mathbf{,}๐ฌ_{๐๐
๐}\mathbf{]}& \mathbf{=}& \mathrm{๐}๐น_{\mathbf{[}๐}^๐๐ฌ_{๐
๐\mathbf{]}๐}^{}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น_๐^๐๐ฌ_{๐๐
๐}\hfill \\ \hfill \mathbf{[}๐ฌ_๐{}_{}{}^{๐}\mathbf{,}๐ฌ_๐{}_{}{}^{๐
}\mathbf{]}& \mathbf{=}& ๐น_๐^๐๐ฌ_๐{}_{}{}^{๐
}\mathbf{}๐น_๐^๐
๐ฌ_๐^๐\hfill & & & \end{array}$$
The elements $`\mathbf{\{}๐^๐\mathbf{,}๐_{๐๐}\mathbf{\}}`$ (or equivalently $`\mathbf{\{}๐_๐\mathbf{,}๐^{๐๐}\mathbf{\}}`$) span the maximal 36-dimensional abelian nilpotent subalgebra of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ .
Finally, the generators are normalized according to
$`\mathrm{๐๐ซ}\mathbf{(}๐ต๐ต\mathbf{)}`$ $`\mathbf{=}`$ $`\mathrm{๐๐}\mathbf{}\mathrm{๐}\mathbf{,}`$
$`\mathrm{๐๐ซ}\mathbf{(}๐^๐๐_๐\mathbf{)}`$ $`\mathbf{=}`$ $`\mathrm{๐๐}๐น_๐^๐\mathbf{,}`$
$`\mathrm{๐๐ซ}\mathbf{(}๐^{๐๐}๐_{๐๐
}\mathbf{)}`$ $`\mathbf{=}`$ $`\mathrm{๐๐}\mathbf{}\mathrm{๐}\mathbf{!}๐น_{๐๐
}^{๐๐}\mathbf{,}`$
$`\mathrm{๐๐ซ}\mathbf{(}๐ฌ_{๐๐๐}๐ฌ^{๐
๐๐}\mathbf{)}`$ $`\mathbf{=}`$ $`\mathrm{๐๐}\mathbf{}\mathrm{๐}\mathbf{!}๐น_{๐๐๐}^{๐
๐๐}\mathbf{,}`$
$`\mathrm{๐๐ซ}\mathbf{(}๐ฌ_๐{}_{}{}^{๐}๐ฌ_{๐}^{}{}_{}{}^{๐
}\mathbf{)}`$ $`\mathbf{=}`$ $`\mathrm{๐๐}๐น_๐^๐
๐น_๐^๐\mathbf{}\frac{\mathrm{๐๐}}{\mathrm{๐}}๐น_๐^๐๐น_๐^๐
\mathbf{,}`$
with all other traces vanishing.
## Appendix Appendix C: Supersymmetry variations of the generalized vielbein
The supersymmetry variations of the achtbein and the relevant components of the 3-form potential with our normalization read as follows in the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ basis:
$`๐น๐_๐^๐`$ $`\mathbf{=}`$ $`\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{\right)}\mathbf{,}`$ (C.1)
$`๐น๐จ_{๐๐๐}`$ $`\mathbf{=}`$ $`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{(}\overline{๐บ}_๐ถ\mathbf{(}๐ช_{\mathbf{[}๐๐}\mathbf{)}_{๐ถ๐ท}๐ฟ_{๐\mathbf{]}}{}_{๐ท}{}^{}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\mathbf{(}\overline{๐ช}_{\mathbf{[}๐๐}\mathbf{)}_{\dot{๐ถ}\dot{๐ท}}๐ฟ_{๐\mathbf{]}}{}_{\dot{๐ท}}{}^{}\mathbf{)}\mathbf{.}`$
(The relative minus signs in the second terms on the r.h.s. are due the fact that the Dirac conjugate spinors are appropriate to $`๐
\mathbf{=}\mathrm{๐}`$ and differ from the ones in $`๐
\mathbf{=}\mathrm{๐๐}`$ by an extra factor $`๐ช^\mathrm{๐}`$). These formulas must now be compared with the $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ covariant ones in terms of the generalized vielbein. The latter are most conveniently computed in terms of the matrices $`\stackrel{\mathbf{~}}{๐ฌ}^๐`$, $`\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}`$ from (3.12), making use of the $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{,}\mathbf{)}`$ decomposition of $`๐ฌ_{\mathrm{๐}\mathbf{(}\mathrm{๐}\mathbf{)}}`$ described in the appendix B. In the special $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ gauge (2.13), (2.15), (2.16) these matrices take the form
$`\stackrel{\mathbf{~}}{๐ฌ}^๐`$ $`\mathbf{=}`$ $`๐^๐{}_{๐}{}^{}๐ง_{}^{๐}\mathbf{=}\mathrm{๐}๐ซ^\mathbf{}\mathrm{๐}๐_๐{}_{}{}^{๐}๐_{}^{๐}\mathbf{,}`$ (C.2)
$`\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}`$ $`\mathbf{=}`$ $`๐_{๐๐๐}๐ง^๐\mathbf{=}\mathrm{๐}๐ซ^\mathbf{}\mathrm{๐}๐_๐{}_{}{}^{๐}๐_{๐}^{}{}_{}{}^{๐}๐_{๐๐}^{}\mathbf{+}๐จ_{๐๐๐}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{,}`$
in the upper Borel subalgebra. The $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ covariant supersymmetry variations have been presented in (2.19) and can be written as
$$๐น\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{=}\mathbf{[}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{,}๐\mathbf{]}\mathbf{,}๐น\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{=}\mathbf{[}\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}\mathbf{,}๐\mathbf{]}\mathbf{,}$$
(C.3)
with $`๐`$ given by
$$๐\mathbf{:=}\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{\left(}๐ช_{๐จ\dot{๐จ}}^๐ฐ\overline{๐บ}^๐ฐ๐^{\dot{๐จ}}๐^๐จ\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{\mathrm{๐๐จ๐ฆ๐ฉ}}^{๐ฐ๐ฑ}๐ฟ^{๐ฐ๐ฑ}\mathbf{\right)}\mathbf{,}$$
(C.4)
where $`๐_{\mathrm{๐๐จ๐ฆ๐ฉ}}^{๐ฐ๐ฑ}๐ฟ^{๐ฐ๐ฑ}`$ is the compensating $`๐บ๐ถ\mathbf{(}\mathrm{๐๐}\mathbf{)}`$ rotation to restore the triangular gauge of (C.2). Upon decomposition into the $`๐บ๐ถ\mathbf{(}\mathrm{๐}\mathbf{)}`$ fields, the first term yields
$`๐ช_{๐จ\dot{๐จ}}^๐ฐ\overline{๐บ}^๐ฐ๐^{\dot{๐จ}}๐^๐จ`$ $`\mathbf{=}`$ $`\mathrm{๐}\mathbf{\left(}\overline{๐บ}_๐ถ๐ฟ_{๐๐ถ}\mathbf{+}\overline{๐บ}_{\dot{๐ถ}}๐ฟ_{๐\dot{๐ถ}}\mathbf{\right)}\mathbf{(}๐^๐\mathbf{+}๐_๐\mathbf{)}`$
$`\mathbf{+}\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ๐ท}^{๐๐}๐ฟ_{๐๐ท}\mathbf{+}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}\dot{๐ท}}^{๐๐}๐ฟ_{๐\dot{๐ท}}\mathbf{\right)}\mathbf{(}๐^๐\mathbf{+}๐_๐\mathbf{)}`$
$`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^{๐๐๐}๐ฟ_{๐\dot{๐ท}}\mathbf{+}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^{๐๐๐}๐ฟ_{๐๐ท}\mathbf{\right)}\mathbf{(}๐_{๐๐}\mathbf{+}๐^{๐๐}\mathbf{)}`$
$`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ๐ท}^{๐๐}๐ฟ_{๐๐ท}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}\dot{๐ท}}^{๐๐}๐ฟ_{๐\dot{๐ท}}\mathbf{\right)}\mathbf{(}๐ฌ^{๐๐๐}\mathbf{+}๐ฌ_{๐๐๐}\mathbf{)}`$
$`\mathbf{}\mathbf{(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{)}\mathbf{\left(}๐ฌ_๐{}_{}{}^{๐}\mathbf{+}๐ฌ_๐{}_{}{}^{๐}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น^{๐๐}๐ต\mathbf{\right)}\mathbf{,}`$
whereas $`๐_{\mathrm{๐๐จ๐ฆ๐ฉ}}^{๐ฐ๐ฑ}๐ฟ^{๐ฐ๐ฑ}`$ is determined in such a way as to rotate $`๐`$ back into the upper Borel subalgebra. The resulting $`๐`$ will then preserve the special gauge choice (C.2). Explicitly, it takes the form
$`๐`$ $`\mathbf{}`$ $`๐ช_{๐จ\dot{๐จ}}^๐ฐ\overline{๐บ}^๐ฐ๐^{\dot{๐จ}}๐^๐จ\mathbf{+}\frac{\mathrm{๐}}{\mathrm{๐}}๐_{\mathrm{๐๐จ๐ฆ๐ฉ}}^{๐ฐ๐ฑ}๐ฟ^{๐ฐ๐ฑ}`$
$`\mathbf{=}`$ $`\mathrm{๐}\mathbf{\left(}\overline{๐บ}_๐ถ๐ฟ_{๐๐ถ}\mathbf{+}\overline{๐บ}_{\dot{๐ถ}}๐ฟ_{๐\dot{๐ถ}}\mathbf{\right)}๐^๐\mathbf{+}\mathrm{๐}\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ๐ท}^{๐๐}๐ฟ_{๐๐ท}\mathbf{+}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}\dot{๐ท}}^{๐๐}๐ฟ_{๐\dot{๐ท}}\mathbf{\right)}๐^๐`$
$`\mathbf{}\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^{๐๐๐}๐ฟ_{๐\dot{๐ท}}\mathbf{+}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^{๐๐๐}๐ฟ_{๐๐ท}\mathbf{\right)}๐_{๐๐}`$
$`\mathbf{}\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ๐ท}^{๐๐}๐ฟ_{๐๐ท}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}\dot{๐ท}}^{๐๐}๐ฟ_{๐\dot{๐ท}}\mathbf{\right)}๐ฌ^{๐๐๐}`$
$`\mathbf{}\mathbf{(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{)}\mathbf{\left(}\mathrm{๐}๐ฌ_๐{}_{}{}^{๐}\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}๐น^{๐๐}๐ต\mathbf{\right)}`$
From (C.3) we then find the supersymmetry variation:
$`๐น\stackrel{\mathbf{~}}{๐ฌ}^๐`$ $`\mathbf{=}`$ $`\mathbf{}\mathrm{๐}๐ซ^\mathbf{}\mathrm{๐}๐^๐\mathbf{\left(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{\right)}`$
$`\mathbf{}\mathrm{๐}๐ซ^\mathbf{}\mathrm{๐}๐^๐๐_๐{}_{}{}^{๐}\mathbf{(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{)}`$
$`๐น\stackrel{\mathbf{~}}{๐ฌ}_{๐๐}`$ $`\mathbf{=}`$ $`\mathrm{๐}๐ซ^\mathbf{}\mathrm{๐}๐_{๐๐}๐_๐{}_{}{}^{๐}\mathbf{(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{)}`$
$`\mathbf{}\mathrm{๐}๐ซ^\mathbf{}\mathrm{๐}๐_{๐๐}๐_๐{}_{}{}^{๐}๐_{๐}^{}{}_{}{}^{๐}\mathbf{(}\overline{๐บ}_๐ถ๐ช_{๐ถ\dot{๐ท}}^๐๐ฟ_{๐\dot{๐ท}}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\overline{๐ช}_{\dot{๐ถ}๐ท}^๐๐ฟ_{๐๐ท}\mathbf{)}`$
$`\mathbf{}\frac{\mathrm{๐}}{\mathrm{๐}}\stackrel{\mathbf{~}}{๐ฌ}^๐\mathbf{(}\overline{๐บ}_๐ถ\mathbf{(}๐ช_{\mathbf{[}๐๐}\mathbf{)}_{๐ถ๐ท}๐ฟ_{๐\mathbf{]}}{}_{๐ท}{}^{}\mathbf{}\overline{๐บ}_{\dot{๐ถ}}\mathbf{(}\overline{๐ช}_{\mathbf{[}๐๐}\mathbf{)}_{\dot{๐ถ}\dot{๐ท}}๐ฟ_{๐\mathbf{]}}{}_{\dot{๐ท}}{}^{}\mathbf{)}`$
$`\mathbf{+}๐จ_{๐๐๐}๐น\stackrel{\mathbf{~}}{๐ฌ}^๐`$
and thus agreement with (C.1), since the one-but-last term is just $`๐น๐จ_{๐๐๐}\stackrel{\mathbf{~}}{๐ฌ}^๐`$.
|
warning/0006/cond-mat0006365.html
|
ar5iv
|
text
|
# c-axis penetration depth in Bi2Sr2CaCu2O8+ฮด single crystals measured by ac-susceptibility and cavity perturbation technique
## Abstract
The $`c`$-axis penetration depth $`\mathrm{\Delta }\lambda _c`$ in Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> (BSCCO) single crystals as a function of temperature has been determined using two techniques, namely, measurements of the ac-susceptibility at a frequency of 100 kHz and the surface impedance at 9.4 GHz. Both techniques yield an almost linear function $`\mathrm{\Delta }\lambda _c(T)T`$ in the temperature range $`T<0.5T_c`$. Electrodynamic analysis of the impedance anisotropy has allowed us to estimate $`\lambda _c(0)50`$ $`\mu `$m in BSCCO crystals overdoped with oxygen ($`T_c84`$ K) and $`\lambda _c(0)150`$ $`\mu `$m at the optimal doping level ($`T_c90`$ K).
Cuprate high-temperature superconductors (HTS) are layered anisotropic materials. Therefore the electrodynamic problem of the magnetic field penetration depth in HTS in the low-field limit is characterized by two length parameters, namely, $`\lambda _{ab}`$ controlled by screening currents running in the CuO<sub>2</sub> planes (in-plane penetration depth) and $`\lambda _c`$ due to currents running in the direction perpendicular to these planes (out-of-plane or $`c`$-axis penetration depth). The temperature dependence of the penetration depth in HTS is largely determined by the superconductivity mechanism. It is known (see, e.g., Ref. and references therein) that $`\mathrm{\Delta }\lambda _{ab}(T)T`$ in the range $`T<T_c/3`$ in high-quality HTS samples at the optimal level of doping, and this observation has found the most simple interpretation in the $`d`$-wave model of the high-frequency response in HTS . Measurements of $`\lambda _c(T)`$ are quoted less frequently than those of $`\lambda _{ab}(T)`$. Most of such data published by far were derived from microwave measurements of the surface impedance of HTS crystals . There is no consensus in literature about $`\mathrm{\Delta }\lambda _c(T)`$ at low temperatures. Even in reports on low-temperature properties of high-quality YBCO crystals, which are the most studied objects, one can find both linear, $`\mathrm{\Delta }\lambda _c(T)T`$ , and quadratic dependences in the range $`T<T_c/3`$. In BSCCO materials, the shape of $`\mathrm{\Delta }\lambda _c(T)`$ depends on the level of oxygen doping: in samples with maximal $`T_c90`$ K $`\mathrm{\Delta }\lambda _c(T)T`$ at low temperatures ; at higher oxygen contents (overdoped samples) $`T_c`$ is lower and the linear function $`\mathrm{\Delta }\lambda _c(T)`$ transforms to a quadratic one . The common feature of all microwave experiments is that the change in the ratio $`\mathrm{\Delta }\lambda _c(T)/\lambda _c(0)`$ is smaller than in $`\mathrm{\Delta }\lambda _{ab}(T)/\lambda _{ab}(0)`$ because in all HTS $`\lambda _c(0)\lambda _{ab}(0)`$. The length $`\lambda _c(0)`$ is especially large in BSCCO crystals, $`\lambda _c(0)>10`$ $`\mu `$m and, according to some estimates, it ranges up to $`500`$ $`\mu `$m. The large spread of $`\lambda _c(0)`$ is caused by two factors, namely, the poor accuracy of the techniques used in determination of $`\lambda _c(0)`$ and effects of local and extended defects in tested samples, whose range is of order of 1 mm and comparable to both $`\lambda _c`$ and total sample dimensions.
Recently we suggested a new technique for determination of $`\lambda _c(0)`$ based on the measurements of the surface barrier field $`H_J(T)1/\lambda _c(T)`$ at which Josephson vortices penetrate into the sample. The field $`H_J`$ corresponds to the onset of microwave absorption in the locked state of BSCCO single crystals. This paper suggests an alternative technique based on comparison between microwave measurements of BSCCO crystals aligned differently with respect to ac magnetic field and a numerical solution of the electrodynamic problem of the magnetic field distribution in an anisotropic plate at an arbitrary temperature. Moreover, since $`\lambda _c(0)`$ in BSCCO single crystals is relatively large, we managed to determine $`\lambda _c(T)`$ from the temperature dependences of ac-susceptibility and compare these measurements to results of microwave experiments.
Single crystals of BSCCO were grown by the floating-zone method and shaped as rectangular platelets. This paper presents measurements of two BSCCO samples with various levels of oxygen doping. The first sample (#1), characterized by a higher critical temperature, $`T_c90`$ K (optimally doped), has dimensions $`a\times b\times c1.5\times 1.5\times 0.1`$ mm<sup>3</sup> ($`ab`$). The second (#2, $`a\times b\times c0.8\times 1.8\times 0.03`$ mm<sup>3</sup>) is slightly overdoped ($`T_c84`$ K).
When measuring the ac-susceptibility $`\chi =\chi ^{}i\chi ^{\prime \prime }`$, we placed a sample inside one of two identical induction coils. The coils were connected to one another, and the out-of-phase and in-phase components of the imbalance signal were measured at a frequency of $`10^5`$ Hz. These components are proportional to the real and imaginary parts of the sample magnetic moment $`M`$, respectively: $`M=\chi vH_0`$, where $`v`$ is the sample volume and $`H_0`$ is the ac magnetic field amplitude, which was within 0.1 Oe in our experiments.
Figure 1 shows temperature dependences $`\chi ^{}(T)/|\chi ^{}(0)|`$ in
sample #1 for three different sample alignments with respect to the ac magnetic field: the transverse (T) orientation, $`๐_\omega ๐`$, (the inset on the left of Fig. 1), when the screening current flows in the $`ab`$-plane (full circles); in the longitudinal (L) orientation, $`๐_\omega ๐`$, (the inset on the right of Fig. 1, $`๐_\omega `$ is parallel to the $`b`$-edge of the crystal), when currents running in the directions of both CuO<sub>2</sub> planes and the $`c`$-axis are present (up triangles); in the L-orientation, $`๐_\omega ๐`$, whose difference from the previous configuration is that the sample is turned around the $`c`$-axis through $`90^{}`$ (down triangles). Fig. 1 clearly shows that at $`T<T_c`$ $`\chi _{ab}^{}(T)`$ is notably smaller in the T-orientation than $`\chi _{ab+c}^{}(T)`$ in the L-orientation (the subscripts of $`\chi ^{}`$ denote the direction of the screening current). The coincidence of $`\chi _{ab+c}^{}(T)`$ curves at $`๐_\omega ๐`$ and the small width of the superconducting transition at $`๐_\omega ๐`$ ($`\mathrm{\Delta }T_c<1`$ K) indicate that the quality of the tested sample #1 is fairly high. This is supported by precision measurements of surface impedance $`Z_s(T)=R_s(T)+iX_s(T)`$ of sample #1 at frequency $`f=9.4`$ GHz in the T-orientation, which are plotted in Fig. 2. The measurement technique was described in detail elsewhere . It applies to both surface impedance components $`R_s(T)`$ and $`X_s(T)`$:
$$R_s=\mathrm{\Gamma }_s\mathrm{\Delta }(1/Q),X_s=2\mathrm{\Gamma }_s\delta f/f,$$
(1)
Here $`\mathrm{\Gamma }_s=\omega \mu _0_VH_\omega ^2๐V/[_SH_s^2๐S]`$ is the sample geometrical factor ($`\omega =2\pi f`$, $`\mu _0=4\pi 10^7`$ H/m, $`V`$ is the volume of the cavity, $`H_\omega `$ is the magnetic field generated in the cavity, $`S`$ is the total sample surface area, and $`H_s`$ is the tangential component of the microwave magnetic field on the sample surface); $`\mathrm{\Delta }(1/Q)`$ is the difference between the values 1/Q of the cavity with the sample inside and empty cavity; $`\delta f`$ is the frequency shift relative to that which would be measured for a sample with perfect screening, i.e., no penetration of the microwave fields. In the experiment we measure the difference $`\mathrm{\Delta }f(T)`$ between resonant frequency shifts with temperature of the loaded and empty cavity, which is equal to $`\mathrm{\Delta }f(T)=\delta f(T)+f_0`$, where $`f_0`$ is a constant . The constant $`f_0`$ includes both the perfect-conductor shift and the uncontrolled contribution caused by opening and closing the cavity. In HTS single crystals, the constant $`f_0`$ can be directly derived from measurements of the surface impedance in the normal state; in particular, in the T-orientation $`f_0`$ can be derived from the condition that the real and imaginary parts of the impedance should be equal above $`T_c`$ (normal skin-effect). In Fig. 2 $`R_s(T)=X_s(T)`$ at $`TT_c`$, and its temperature dependence is adequately described by the expression $`2R_s^2(T)/\omega \mu _0=\rho (T)=\rho _0+bT`$ with $`\rho _013`$ $`\mu \mathrm{\Omega }`$cm and $`b0.3`$ $`\mu \mathrm{\Omega }`$cm/K. Given $`R_s(T_\mathrm{c})=\sqrt{\omega \mu _0\rho (T_\mathrm{c})/2}0.12`$ $`\mathrm{\Omega }`$, we obtain the resistivity $`\rho (T_\mathrm{c})40`$ $`\mu \mathrm{\Omega }`$cm. The insets to Fig. 2 show $`R_s(T)`$ and $`\lambda (T)=X_s(T)/\omega \mu _0`$ for $`T<0.7T_\mathrm{c}`$ plotted on a linear scale. The extrapolation of the low-temperature sections of these curves to $`T=0`$ yields estimates of $`\lambda _{ab}(0)2600`$ ร
and the residual surface resistance $`R_{\mathrm{res}}0.5`$ m$`\mathrm{\Omega }`$. $`R_{\mathrm{res}}`$ is due to various defects in the surface layer of the superconductor and it is generally accepted that the lower the $`R_{\mathrm{res}}`$, the better the sample quality. The above mentioned parameters of sample #1 indicate that its quality is fairly high. In the T-orientation, linear functions $`\mathrm{\Delta }R_s(T)`$ and $`\mathrm{\Delta }\lambda _{ab}(T)`$ in the low-temperature range were previously observed in optimally doped BSCCO crystals at a frequency of about 10 GHz . In the slightly overdoped sample #2 we also observed $`\mathrm{\Delta }\lambda _{ab}(T)`$, $`\mathrm{\Delta }R_s(T)T`$ at low temperature, moreover, the measurement $`R_{\mathrm{res}}120`$ $`\mu \mathrm{\Omega }`$ is, to the best of our knowledge, the lowest value ever obtained in BSCCO single crystals.
In both superconducting and normal states of HTS, the relation between electric field and current density is local: $`j=\widehat{\sigma }E`$, where the conductivity $`\widehat{\sigma }`$ is a tensor characterized by components $`\sigma _{ab}`$ and $`\sigma _c`$. In the normal state, ac field penetrates in the direction of the $`c`$-axis through the skin depth $`\delta _{ab}=\sqrt{2/\omega \mu _0\sigma _{ab}}`$ and in the CuO<sub>2</sub> plane through $`\delta _c=\sqrt{2/\omega \mu _0\sigma _c}`$. In the superconducting state all parameters $`\delta _{ab}`$, $`\delta _c`$, $`\sigma _{ab}=\sigma _{ab}^{}i\sigma _{ab}^{\prime \prime }`$, and $`\sigma _c=\sigma _c^{}i\sigma _c^{\prime \prime }`$ are complex. In the temperature range $`T<T_c`$, if $`\sigma ^{}\sigma ^{\prime \prime }`$, the field penetration depths are given by the formulas $`\lambda _{ab}=\sqrt{1/\omega \mu _0\sigma _{ab}^{\prime \prime }}`$, $`\lambda _c=\sqrt{1/\omega \mu _0\sigma _c^{\prime \prime }}`$. In the close neighborhood of $`T_c`$, if $`\sigma ^{}\sigma ^{\prime \prime }`$, the decay of magnetic field in the superconductor is characterized by the functions $`\mathrm{Re}(\delta _{\mathrm{ab}})`$ and $`\mathrm{Re}(\delta _\mathrm{c})`$, which turn to $`\delta _{ab}`$ and $`\delta _c`$, respectively, at $`TT_c`$.
In the L-orientation of BSCCO single crystals at $`T<0.9T_c`$ the penetration depth is smaller than characteristic sample dimensions. If we neglect the anisotropy in the $`ab`$-plane and the contribution from $`ac`$-faces (see the inset to Fig. 1), which is a factor $`c/b`$ smaller than that of the $`ab`$-surfaces, the effective impedance $`Z_s^{ab+c}`$ in the L-orientation can be expressed in terms of $`Z_s^{ab}`$ and $`Z_s^c`$ averaged over the surface area (the superscripts of $`Z_s`$ denote the direction of the screening current). Thus, given measurements of $`\mathrm{\Delta }\lambda _{ab}(T)=\mathrm{\Delta }X_s^{ab}(T)/\omega \mu _0`$ in the T-orientation and of the effective value $`\mathrm{\Delta }\lambda _{ab+c}(T)=\mathrm{\Delta }X_s^{ab+c}(T)/\omega \mu _0`$ in the L-orientation, we obtain
$$\mathrm{\Delta }\lambda _c=\left[(a+c)\mathrm{\Delta }\lambda _{ab+c}a\mathrm{\Delta }\lambda _{ab}\right]/c.$$
(2)
This technique for determination of $`\mathrm{\Delta }\lambda _c(T)`$ was used in microwave experiments at low temperatures, $`T<T_c`$. Even so, it cannot be applied to the range of higher temperatures because the size effect plays an important role. Really, at $`T>0.9T_\mathrm{c}`$ the lengths $`\lambda _c`$ and $`\delta _c`$ are comparable to the sample dimensions. In order to analyze our measurements in both superconducting and normal states of BSCCO, we used formulas for field distributions in an anisotropic long strip ($`ba,c`$) in the L-orientation. These formulas neglect the effect of $`ac`$-faces of the crystal, but take account of the size effect. In addition, in a sample shaped as a long strip, there is a simple relation between its surface impedance components and ac-susceptibility, which is expressed in terms of parameter $`\mu `$ introduced in Ref. :
$$\mathrm{\Delta }(1/Q)2i\delta f/f=i\gamma \mu v/V,\chi =1+\mu ,$$
(3)
where $`\gamma =VH_0^2/[_VH_\omega ^2๐V]=10.6`$ is a constant characterizing our cavity . At an arbitrary temperature, the complex parameter $`\mu =\mu ^{}i\mu ^{\prime \prime }`$ is controlled by the components $`\sigma _{ab}(T)`$ and $`\sigma _c(T)`$ of the conductivity tensor:
$$\mu =\frac{8}{\pi ^2}\underset{n}{}\frac{1}{n^2}\left\{\frac{\mathrm{tan}(\alpha _n)}{\alpha _n}+\frac{\mathrm{tan}(\beta _n)}{\beta _n}\right\},$$
(4)
where the sum is performed over odd integers $`n>0`$, and
$`\alpha _n^2={\displaystyle \frac{a^2}{\delta _c^2}}\left({\displaystyle \frac{i}{2}}+{\displaystyle \frac{\pi ^2}{4}}{\displaystyle \frac{\delta _{ab}^2}{c^2}}n^2\right),\beta _n^2={\displaystyle \frac{c^2}{\delta _{ab}^2}}\left({\displaystyle \frac{i}{2}}+{\displaystyle \frac{\pi ^2}{4}}{\displaystyle \frac{\delta _c^2}{a^2}}n^2\right).`$ (5)
In the superconducting state at $`T<0.9T_c`$ we find that $`\lambda _{ab}c`$ and $`\lambda _ca`$. In this case, we derive from Eq. (4) a simple expression for the real part of $`\mu `$:
$$\mu ^{}=1+\chi ^{}=\frac{2\lambda _c}{a}+\frac{2\lambda _{ab}}{c}.$$
(6)
One can easily check up that in the range of low temperatures the change in $`\mathrm{\Delta }\lambda _\mathrm{c}(T)`$ prescribed by Eq. (5) is identical to Eq. (2). Figure 3 shows measurements of $`\mathrm{\Delta }\lambda _c(T)`$ in sample #1 (circles) and sample #2 (squares) at $`T<0.9T_c`$. The
open symbols plot low-frequency measurements obtained in accordance with Eq. (5), the full symbols plot microwave measurements processed by Eq. (2). Agreement between measurements of sample #2 (lower curve) is fairly good, but in fitting together experimental data from sample #1 (upper curve) we had to divide by a factor of 1.8 all $`\mathrm{\Delta }\lambda _c(T)`$ derived from measurements of ac-susceptibility using Eq. (5). The cause of the difference between $`\mathrm{\Delta }\lambda _c(T)`$ measured in sample #1 at different frequencies is not quite clear. We rule out a systematic experimental error that could be caused by misalignment of the sample with respect to the ac magnetic field in the coil because (i) curves of $`\mathrm{\Delta }\lambda _c(T)`$ were accurately reproduced when square sample #1 was turned through an angle of $`90^{}`$ in the L-orientation, and (ii) a small tilt of this sample with respect to the magnetic field generated by the coil would lead to a larger difference (more than a factor of 1.8) between the two sets of experimental data. It seems more plausible that Eqs. (2) and (5), which neglect the contribution of $`ac`$-faces, yield inaccurate results concerning sample #1: its $`ac`$-faces, which have a notable area (sample #1 is thick), can host a lot of defects (for example, those of the capacitive type), and the latter can affect the character of field penetration as a function of frequency.
The curves of $`\mathrm{\Delta }\lambda _c(T)`$ at $`T<0.5T_c`$ plotted in Fig. 3 are almost linear: $`\mathrm{\Delta }\lambda _c(T)T`$. The inset to Fig. 3 shows the low-temperature section of the curve of $`\mathrm{\Delta }\lambda _c(T)`$ in sample #1. Its slope is 0.3 $`\mu `$m/K and equals that from Ref. . Note also that changes in $`\mathrm{\Delta }\lambda _c(T)`$ are smaller in the oxygen-overdoped sample #2 than in sample #1.
We can estimate $`\lambda _c(0)`$ by substituting in Eq. (1) $`\delta f(0)`$ obtained by comparing of $`\mathrm{\Delta }(1/Q)`$ and $`\mathrm{\Delta }f=\delta ff_0`$ measurements taken in the T- and L-orientations to numerical calculations by Eqs. (3) and (4), which take account of the size effect in the high-frequency response of an anisotropic crystal. The procedure of comparison for sample #1 is illustrated by Fig. 4. Unlike the case of the T-orientation, the measured temperature dependence of $`\mathrm{\Delta }(1/Q)`$
in the L-orientation deviates from $`(2\mathrm{\Delta }f/f)`$ owing to the size effect. Using the measurements of $`R_s=\sqrt{\omega \mu _0/2\sigma _{ab}}`$ at $`T>T_c`$ in the T-orientation (Fig. 2) for determination of $`\sigma _{ab}(T)`$, alongside the data on $`\mathrm{\Delta }(1/Q)`$ in the L-orientation (open squares in Fig. 4), from Eqs. (3) and (4) we obtain the curve of $`\rho _c(T)=1/\sigma _c(T)`$ shown in the right-hand inset to Fig. 4. Further, using the functions $`\sigma _c(T)`$ and $`\sigma _{ab}(T)`$, we calculate $`(2\delta f/f)`$ versus temperature for $`T>T_c`$, which is plotted by the solid line in Fig. 4. This line is approximately parallel to the experimental curve of $`2\mathrm{\Delta }f/f`$ in the L-orientation (open circles in Fig. 4). The difference $`2(\delta f\mathrm{\Delta }f)/f`$ yields the additive constant $`f_0`$. Given $`f_0`$ and $`\mathrm{\Delta }f(T)`$ measured in the range $`T<T_c`$, we also obtain $`\delta f(T)`$ in the superconducting state in the L-orientation. As a result, with due account of $`\lambda _{ab}(T)`$ (the inset to Fig. 2), we derive from Eqs. (3) and (5) $`\lambda _c(0)`$, which equals approximately 150 $`\mu `$m in sample #1. A similar procedure performed with sample #2 yields $`\lambda _c(0)50`$ $`\mu `$m, which is in agreement with our measurements of overdoped BSCCO obtained using a different technique . We also estimated $`\lambda _c(0)`$ on the base of absolute measurements of the susceptibility $`\chi _c^{}(0)`$ from Eq. (5), and we obtained $`\lambda _c(0)210`$ $`\mu `$m for sample #1 and $`\lambda _c(0)70`$ $`\mu `$m for sample #2. These results are in reasonable agreement with our microwave measurements if we take into consideration the fact that the accuracy of $`\lambda _c`$ measurements is rather poor and the error can be up to 30%.
In conclusion, we have used the ac-susceptibility and cavity perturbation techniques in studying anisotropic high frequency properties of BSCCO single crystals. We have observed almost linear dependences $`\mathrm{\Delta }\lambda _c(T)T`$, which are in fair agreement with both experimental and theoretical results by other researchers. We have also investigated a new technique for determination of $`\lambda _c(0)`$, which is a factor of three higher in the optimally doped BSCCO sample than in the overdoped crystal. The ratio between the slopes of curves of $`\mathrm{\Delta }\lambda _c(T)`$ in the range $`TT_c`$ is the same. These facts could be put down to dependences of $`\lambda _c(0)`$ and $`\mathrm{\Delta }\lambda _c(T)`$ on the oxygen content in these samples. At the same time, we cannot rule out influence of defects in the samples on $`\lambda _c(0)`$, even though their quality in the $`ab`$-plane is fairly high, according to our experiments. In order to draw ultimate conclusions concerning the nature of the transport properties along the $`c`$-axis in BSCCO single crystals, studies of more samples with various oxygen contents are needed.
We would like to thank V. F. Gantmakher for helpful discussions. This research was supported by grant No. 4985 of CNRS-RAS cooperation. The work at ISSP was also supported by the Russian Fund for Basic Research (grants 97-02-16836 and 98-02-16636) and Scientific Council on Superconductivity (project 96060).
|
warning/0006/astro-ph0006128.html
|
ar5iv
|
text
|
# High-Resolution Spectroscopy from 3050 to 10000 ร
of the HDF-S QSO J2233-606 with UVES at the ESO VLT 1footnote 11footnote 1Based on material collected with the ESO VLT telescope
## 1 Introduction
Starting on September 28, 1998 and for two weeks, the Hubble Space Telescope aimed at the same narrow slice of sky in the constellation Tucana. The observing strategy of the Hubble Deep Field South (HDFโS, Williams et al., (2000)) differs from its northern analogous in several respects. The Space Telescope Imaging Spectrograph field was centered on a relatively bright ($`B17.5`$) quasar at intermediate redshift (J2233-606, $`z_{\mathrm{em}}=2.238`$). The installation of STIS and NICMOS on HST in 1997 February has enabled parallel observations with three cameras. In this way the HDFโS dataset includes deep WFPC2 imaging (Casertano et al., (2000)), UV-Visible imaging (Gardner et al., (2000)) and spectroscopy (Savaglio et al., (1999)), deep near-infrared imaging (Fruchter et al., (2000)), and wider-area flanking field observations (Lucas et al., (2000)). Ground-based observations have been carried out at ESO with the VLT and NTT telescopes in the framework of the ESO HDF-South Project (http://www.eso.org/paranal/sv/svhdfs.html). The simultaneous availability of deep imaging and a large spectroscopic coverage at medium-high resolution makes the HDFโS a unique field to study the relationship between galaxies and absorbers, the quasar environment, the abundance pattern of metal absorption systems.
Spectroscopic observations of J2233-606 have already been reported by Savaglio (1998, SAV98), Sealey et al. (1998), Outram et al. (1999, OUT99), and Savaglio et al. (1999). The relationship between the low redshift ($`z1`$) galaxies lying within 1โ of the QSO and the absorption systems is described by Tresse et al. (1999), who also found another QSO in the field at an angular separation of 44.5โ from the HDF-S QSO, with $`z_{\mathrm{em}}=1.336`$.
In this paper we present high-resolution spectroscopy ($`\mathrm{}45000`$) carried out with the VLT UV-Visual Echelle Spectrograph, UVES (Dekker et al., (2000)). These new data are unique in terms of $`S/N`$ and spectral range (3050-10000 ร
) and ideally match the HST STIS observations obtained in the interval 2275-3118 ร
with a FWHM resolution of 10 km s<sup>-1</sup> (Savaglio et al., (1999)). In Sect. 2 the observations and data reduction are described. Sect. 3 deals with the process of identification of the absorption lines and their fit through $`\chi ^2`$ minimization of Voigt profiles. In Sect. 4 the general properties of the Ly$`\alpha `$ forest are discussed, while Sect. 5 describes the 24 identified metal systems.
## 2 The Observations
The UVES observation of J2233-606 were carried out during the commissioning of the instrument in October 1999. Details are given in Table LABEL:tab:obs. Two set-ups were used: the first (dichroic 1) covering simultaneously the 3050-3875 ร
and 4795-6815 ร
ranges in the blue and red arm, respectively; the second (dichroic 2) covering simultaneously the 3760-5004 ร
and 6725-10000 ร
ranges in the blue and red arm, respectively.
The data have been reduced in the new UVES context available in the 99NOV edition of MIDAS, the ESO data reduction system. For the wavelength calibration Thorium lamp spectra have been used. Wavelengths have then been corrected to vacuum heliocentric values.
The final combined spectrum has been rebinned to a constant pixel size of 0.05 ร
and covers the wavelength range $`\lambda \lambda \mathrm{\hspace{0.17em}3050}10000`$ ร
.
The resolution, as measured from the Thorium lines of the calibration spectra, extracted and treated in the same way as the QSOs spectra, is $`\mathrm{}45000`$. The $`S/N`$ ratio of the final spectrum is about 50 per resolution element at 4000 ร
, 90 at 5000 ร
, 80 at 6000 ร
, 40 at 8000 ร
.
The continuum level was established selecting regions apparently free of absorptions and then fitting a cubic spline with a large smoothness parameter. The normalized spectra were then spliced together to form a total spectrum.
A section of the normalized spectrum from 3100 to 4300 ร
is shown in Fig. 1-2. The full spectrum is available in the form of FITS tables at the URL http://www.stecf.org/hstprogrammes/J22/J22.html.
## 3 Line identification and profile fitting
The MIDAS package FITLYMAN (Fontana & Ballester 1995) has been used to measure the parameters of the absorption lines. Line fitting through $`\chi ^2`$ minimization of Voigt profiles has been carried out in order to determine the redshifts, column densities and Doppler widths of the identified features. When possible, we based our fit on the components found for multiple ions at the same redshift. The absorption-line parameter fits are presented in Table 2. Heavy element systems are identified, in general, by the presence of the C IV and/or Mg II doublets. The spectral separation and intensity ratio between the lines of the doublet allow a firm identification. Subsequently, the strongest ions commonly observed in quasar spectra are searched for at the same redshift. The atomic parameters for the absorption lines observed in QSO spectra have been taken from Morton et al. (1991) and Verner, Barthel & Tytler (1994). All the lines not identified as metals in the Ly$`\alpha `$ forest are fitted as H I Ly$`\alpha `$ and Ly$`\beta `$.
If a strong ion is not detected in the spectrum, we report only an upper limit to the column density. The estimate is based on the fact that there is a linear relation between rest equivalent width and column density for weak lines (because they lie on the linear part of the curve of growth). First, the equivalent width limit for that region is computed and the corresponding column density is given by:
$$N_{\mathrm{lim}}(\mathrm{cm}^2)1.13\times 10^{20}w_{\mathrm{lim}}(\text{ร
})/(\lambda _{\mathrm{rest}}^2(\text{ร
})\times f_{\mathrm{osc}}).$$
(1)
## 4 The Ly$`\alpha `$ Forest
A total of 186 Ly$`\alpha `$ lines are observed in the interval $`\lambda \lambda \mathrm{\hspace{0.17em}3321.3}3885`$ ร
, i.e. between the onset of the Ly$`\beta `$ forest and the range affected by the proximity effect. The HI column density distribution for $`13<\mathrm{log}N(`$H I$`)<17.5`$ is consistent with a power-law distribution with a slope $`1.41\pm 0.05`$, slightly flatter but not significantly different from the canonical value observed at higher redshift (Giallongo et al. (1996)). The Doppler parameter distribution peaks between 20-25 km s<sup>-1</sup>, with a lower cutoff at 15 km s<sup>-1</sup>for lines with $`\mathrm{log}N(`$H I$`)>13`$. Fig.3 shows the number density of Ly$`\alpha `$ lines with $`\mathrm{log}N(`$H I$`)>14`$ observed in the spectrum of J2233-606. At a medium redshift $`<z>=1.96\pm 0.23`$ we observe 33 such lines, 14 of which are associated to metal systems (see below). The corresponding $`dn/dz`$ is $`71\pm 12`$ including metal systems and $`41\pm 9`$ excluding metal systems. The previous estimate of the $`dn/dz`$ in J2233-606 by Savaglio et al. (1999) is in good agreement with the present result. The comparison with data in the literature from other sight-lines, provides a consistent picture of the evolution of the density of Ly$`\alpha `$ absorbers with redshift. In particular, the measurement derived from the spectrum of J2233-606 is remarkably close to the HST point at $`<z>=1.6`$ by Weymann et al. (1998), which is based on 19 lines observed in the spectrum of the QSO UM 18. Fig. 3 suggests that a sharp upturn in the density of Ly$`\alpha `$ lines with $`\mathrm{log}N(`$H I$`)>14`$ takes place at a redshift $`z=1.41.6`$.
The density of Ly$`\alpha `$ lines in J2233-606 not associated with metal systems (see next Section) is also shown in Fig. 3. The agreement with the corresponding data of Giallongo et al. (1996) is very good <sup>2</sup><sup>2</sup>2 the definition of โLy$`\alpha `$ line not associated with metal systemsโ is dependent on the data quality and is therefore not rigorous..
## 5 Metal systems
In general, we consider a group of absorption lines due to the same ion as an individual metal system when the components are blended or not separated by more than a few tens of km s<sup>-1</sup>.
The strongest systems in the spectrum of J2233-606 were already detected by OUT99 and by SAV98. We started from their detections and looked for weaker components and more ions belonging to the same systems. Besides, we found 4 new Mg II doublets, 7 new C IV doublets and a new associated system showing C IV and N V absorptions.
In the following, we list the systems and give a short description of their main features. The quoted redshift is the one of the strongest component or the average redshift in the case of complex systems (i.e. more than three components).
All the metal systems are resumed in Table LABEL:tab:metsys.
### 5.1 Intervening systems
#### 5.1.1 System at $`z_{\mathrm{abs}}=0.000002`$
Strong Na I $`\lambda \lambda \mathrm{\hspace{0.17em}5891},5897`$ is seen from the interstellar medium of the Galaxy.
#### 5.1.2 System at $`z_{\mathrm{abs}}=0.0002`$
The complex Ca II $`\lambda \lambda \mathrm{\hspace{0.17em}3934},3969`$ at this redshift was already observed by OUT99. We detected two more, weaker components. The $`\lambda \mathrm{\hspace{0.17em}3969}`$ line is blended with the N V $`\lambda \mathrm{\hspace{0.17em}1238}`$ feature at $`z_{\mathrm{abs}}=2.206`$.
#### 5.1.3 System at $`z_{\mathrm{abs}}=0.41426`$
A single Mg II $`\lambda \lambda \mathrm{\hspace{0.17em}2796},2803`$ doublet occurs at this redshift (OUT99). The fit with one component yields a relatively poor result, since the equivalent width ratio of the first line, Mg II $`\lambda \mathrm{\hspace{0.17em}2796}`$, relative to the Mg II $`\lambda \mathrm{\hspace{0.17em}2803}`$ line is lower than expected. A fit with a double structure does not give a better result.
Tresse et al. (1999) observed three galaxies at $`z0.4147`$ in the 1โ field around the quasar. They claim that another object, closer to the quasar and belonging to the same over-density of galaxies, could be the responsible for this absorption system.
#### 5.1.4 System at $`z_{\mathrm{abs}}=0.57017`$
Tresse et al. (1999) found at a redshift $`z=0.57`$ a late-type spiral galaxy at an impact parameter of $`18h^1`$ kpc<sup>3</sup><sup>3</sup>3 We consider $`h=H_0/100`$ km/s/Mpc and $`q_0=0.5`$. H I absorption likely associated with this galaxy is seen in the Lyman series, the fit of the Lyman limit in the STIS spectrum gives $`N(`$H I$`)10^{16.8}`$ cm<sup>-2</sup>.
The UVES spectrum shows the associated Mg II doublet, with the strongest component at $`z_{\mathrm{abs}}=0.57017`$. We fitted the system with three components, spanning $`86`$ km s<sup>-1</sup>.
#### 5.1.5 System at $`z_{\mathrm{abs}}=0.58008`$
This new, two component Mg II doublet, seen in the UVES spectrum, falls at the redshift of another galaxy found in the field by Tresse et al. (1999) at an impact parameter $`112h^1`$ kpc from the quasar line of sight.
#### 5.1.6 System at $`z_{\mathrm{abs}}=0.64498`$
A galaxy at a redshift $`z=0.6465`$ was observed at an impact parameter $`220h^1`$ kpc from the quasar line of sight by Tresse et al. (1999). The UVES spectrum shows a single Mg II doublet at this slightly lower absorption redshift.
#### 5.1.7 System at $`z_{\mathrm{abs}}=0.75313`$
A new Mg II doublet was found at this redshift in the UVES spectrum.
#### 5.1.8 System at $`z_{\mathrm{abs}}=1.09169`$
New C IV $`\lambda \lambda \mathrm{\hspace{0.17em}1548},1550`$ system, fitted with two components.
#### 5.1.9 System at $`z_{\mathrm{abs}}=1.32153`$
A new C IV doublet in the Ly$`\alpha `$ forest is observed at this redshift. There is one, clear, unblended component in both $`\lambda \mathrm{\hspace{0.17em}1548}`$ and $`\lambda \mathrm{\hspace{0.17em}1550}`$ lines. On the basis of the absence of the corresponding Lyman $`\beta `$ line, we assumed that the whole feature is C IV $`\lambda \mathrm{\hspace{0.17em}1548}`$ and not H I Ly$`\alpha `$ and we fitted it with three components. We tentatively identified a feature showing the same velocity profile as C IV to be the Si IV $`\lambda \mathrm{\hspace{0.17em}1393}`$ absorption, while the corresponding Si IV $`\lambda \mathrm{\hspace{0.17em}1402}`$ line is blended with a strong H I Lya complex. At a slightly higher redshift, a weak Mg II doublet has been observed showing two stronger components and a flat one in Mg II $`\lambda \mathrm{\hspace{0.17em}2796}`$. At this redshift, also the lines of Fe II $`\lambda \mathrm{\hspace{0.17em}2600}`$, Fe II $`\lambda \mathrm{\hspace{0.17em}2382}`$, C II $`\lambda \mathrm{\hspace{0.17em}1334}`$, Si II $`\lambda \mathrm{\hspace{0.17em}1526}`$ have been detected.
#### 5.1.10 System at $`z_{\mathrm{abs}}=1.3368`$
Tresse et al. (1999) found another quasar in the field at a redshift $`z1.3360`$ and at an angular separation of 44.5โ from the current object. They measured the H I column density from the Ly$`\alpha `$ and Ly$`\beta `$ absorption lines, observed in the high resolution portion of the STIS spectrum. A one component fit gives $`N(`$H I$`)\mathrm{5\; 10}^{15}`$ cm<sup>-2</sup> and a Doppler parameter $`b50`$ km s<sup>-1</sup>.
The UVES spectrum reveals a complex C IV doublet absorption at a slightly higher redshift, covering almost 200 km s<sup>-1</sup>. Even though the system falls in the Ly$`\alpha `$ forest, the identification is quite reliable due to the characteristic profile. The C IV $`\lambda \mathrm{\hspace{0.17em}1548}`$ feature is blended with the possible Si II $`\lambda \mathrm{\hspace{0.17em}1260}`$ of the system at $`z_{\mathrm{abs}}=1.87`$; while the reddest component of C IV $`\lambda \mathrm{\hspace{0.17em}1550}`$ could be blended with the possible N V $`\lambda \mathrm{\hspace{0.17em}1238}`$ at $`z_{\mathrm{abs}}=1.926`$. We detected also a weak, less extended Mg II doublet on the lower redshift side of the C IV absorption and a weak, possible Si IV doublet corresponding to the highest redshift component of C IV. Finally, a fit of the H I Ly$`\alpha `$ line in the STIS spectrum with the same velocity structure as the C IV system gives a total column density of $`N(`$H I$`)\mathrm{3.7\; 10}^{16}`$ cm<sup>-2</sup>.
#### 5.1.11 System at $`z_{\mathrm{abs}}=1.4831`$
A complex of five C IV $`\lambda \lambda \mathrm{\hspace{0.17em}1548},1550`$ doublets, spanning 75 km s<sup>-1</sup>, is seen within the Ly$`\alpha `$ forest. The identification (by OUT99) is firm thanks to the distinctive pattern of the absorption. No other associated metal line is identified.
#### 5.1.12 System at $`z_{\mathrm{abs}}=1.5034`$
We cannot confirm the existence of this system tentatively identified by SAV98.
The $`3\sigma `$ column density limit on the Mg II $`\lambda \mathrm{\hspace{0.17em}2796}`$ absorption is $`N(`$Mg II$`)\stackrel{<}{}1.6\times 10^{11}`$ cm<sup>-2</sup>. Other strong low ionization lines like Mg I $`\lambda \mathrm{\hspace{0.17em}2852}`$ and Fe II $`\lambda \mathrm{\hspace{0.17em}2382}`$ are not observed and there is not the expected strong Ly$`\alpha `$ line in the STIS spectrum.
#### 5.1.13 System at $`z_{\mathrm{abs}}=1.55635`$
A new, possible, weak C IV doublet has been found at this redshift. The corresponding Ly$`\alpha `$ line has an unexpectedly low column density, $`N(`$H I$`)8\times 10^{13}`$ cm<sup>-2</sup>.
#### 5.1.14 System at $`z_{\mathrm{abs}}=1.59055`$
This two component C IV doublet, separated by 9 km s<sup>-1</sup>, was already observed by OUT99.
No Si IV absorption was detected for this system; $`N(`$Si IV$`)\stackrel{<}{}4.3\times 10^{11}`$ cm<sup>-2</sup> is a $`3\sigma `$ upper limit on the column density. The UVES spectrum shows a possible N V doublet feature at the redshift of the stronger component. We fitted the corresponding Ly$`\alpha `$ line with the same components as the C IV doublet.
#### 5.1.15 System at $`z_{\mathrm{abs}}=1.73165`$
A possible, weak C IV doublet has been found in the UVES spectrum, coinciding with a $`N(`$H I$`)1.2\times 10^{14}`$ cm<sup>-2</sup> Ly$`\alpha `$ line.
#### 5.1.16 System at $`z_{\mathrm{abs}}=1.78618`$
This system was already reported by Sealey et al. (1998). C IV has been observed by SAV98 and by OUT99.
The UVES spectrum does not confirm the presence of Mg I $`\lambda \mathrm{\hspace{0.17em}2852}`$ tentatively detected by SAV98. The $`3\sigma `$ limit on this ion is $`N(`$Mg I$`)\stackrel{<}{}7.2\times 10^{10}`$ cm<sup>-2</sup>. On the other hand, the Si IV $`\lambda \lambda \mathrm{\hspace{0.17em}1393},1402`$ is visible in the Ly$`\alpha `$ forest, although there is clear blending with H I lines. The corresponding H I Ly$`\alpha `$ line was fitted with two components for a total column density, $`N(`$H I$`)3.5\times 10^{15}`$ cm<sup>-2</sup>.
#### 5.1.17 System at $`z_{\mathrm{abs}}=1.81565`$
At this redshift, a possible, extremely weak C IV doublet absorption is observed, the $`\lambda \mathrm{\hspace{0.17em}1550}`$ line is barely visible. We fitted the feature with two components, together with the corresponding H I Ly$`\alpha `$ line whose total column density is $`N(`$H I$`)1.4\times 10^{15}`$ cm<sup>-2</sup>.
#### 5.1.18 System at $`z_{\mathrm{abs}}=1.8693`$
This complex heavy element system was already reported by OUT99 and by SAV98.
The UVES spectrum, besides showing the Si IV $`\lambda \lambda \mathrm{\hspace{0.17em}1393},1402`$ doublet, partially resolves the C IV $`\lambda \lambda \mathrm{\hspace{0.17em}1548},1550`$ features and shows the associated absorption by Si III $`\lambda \mathrm{\hspace{0.17em}1206}`$. N V $`\lambda \mathrm{\hspace{0.17em}1238}`$ is blended with a H I Ly$`\alpha `$, the $`3\sigma `$ limit on N V from the $`\lambda \mathrm{\hspace{0.17em}1242}`$ line is: $`N\stackrel{<}{}3.4\times 10^{12}`$ cm<sup>-2</sup>. As for the low ionization elements, the Mg II doublet is detected in correspondence of the high redshift components of Si IV and C II $`\lambda \mathrm{\hspace{0.17em}1334}`$, Si II $`\lambda \lambda \mathrm{\hspace{0.17em}1190},1193`$, $`\lambda \mathrm{\hspace{0.17em}1526}`$, $`\lambda \mathrm{\hspace{0.17em}1260}`$, Al II $`\lambda \mathrm{\hspace{0.17em}1670}`$ absorption are also present. $`N(`$Fe II$`)\stackrel{<}{}4.3\times 10^{11}`$ cm<sup>-2</sup> is a $`3\sigma `$ limit on the column density of Fe II $`\lambda \mathrm{\hspace{0.17em}2382}`$. The H I Lya line falls in the wavelength range of the UVES spectrum, the feature is heavily saturated. We obtained an acceptable fit with three components, even if the large values of the Doppler parameters and the velocity profiles of the metal absorptions suggest a more complex structure. We defer the analysis and discussion of this interesting system to a following paper (DโOdorico and Petitjean, 2000).
The Mg I $`\lambda \mathrm{\hspace{0.17em}2853}`$ absorption tentatively detected by SAV98, has been identified with a telluric line by comparing the spectrum with that of a standard star obtained at comparable resolution.
#### 5.1.19 System at $`z_{\mathrm{abs}}=1.92599`$
A prominent Ly$`\alpha `$ was observed at this redshift in the UCLES spectrum (Outram et al., (1999)).
In the UVES spectrum the C IV $`\lambda \lambda \mathrm{\hspace{0.17em}1548},1550`$ doublet is resolved. We fitted together C IV, Si IV $`\lambda \lambda \mathrm{\hspace{0.17em}1393},1402`$, and Si III $`\lambda \mathrm{\hspace{0.17em}1206}`$, with three components, spanning $`60`$ km s<sup>-1</sup>. The N V $`\lambda \mathrm{\hspace{0.17em}1238}`$ is blended with the C IV $`\lambda \mathrm{\hspace{0.17em}1550}`$ of the system at $`z_{\mathrm{abs}}=1.337`$. The possible weaker component is visible, while the existence of the stronger one is ruled out by the absence of the corresponding N V $`\lambda \mathrm{\hspace{0.17em}1242}`$ line. As for the low ionization lines, they show a shift of -4 km s<sup>-1</sup> with respect to the strongest component of the high ionization absorptions. C II $`\lambda \mathrm{\hspace{0.17em}1334}`$ has been fitted with two components, while for Si II $`\lambda \mathrm{\hspace{0.17em}1260}`$ and a possible Mg II doublet, we fitted only the strongest one. The $`3\sigma `$ limit on the abundance of Fe II based on the absorption at $`\lambda \mathrm{\hspace{0.17em}2382}`$ is: $`N(`$Fe II$`)\stackrel{<}{}4.9\times 10^{11}`$ cm<sup>-2</sup>.
A simultaneous fit of H I Ly$`\alpha `$, Ly$`\beta `$, Ly$`ฯต`$ and Ly8 has been carried out.
#### 5.1.20 System at $`z_{\mathrm{abs}}=1.9422`$
This system presents many ions of different elements at different ionization stages. OUT99 and SAV98 already detected the absorption due to C II $`\lambda \mathrm{\hspace{0.17em}1334}`$, Si II $`\lambda \lambda \mathrm{\hspace{0.17em}1260},1304`$, Si III $`\lambda \mathrm{\hspace{0.17em}1206}`$, and Si IV $`\lambda \lambda \mathrm{\hspace{0.17em}1393},1402`$, C IV $`\lambda \lambda \mathrm{\hspace{0.17em}1548},1550`$, Mg II $`\lambda \lambda \mathrm{\hspace{0.17em}2796},2803`$, Al III $`\lambda \lambda \mathrm{\hspace{0.17em}1854},1862`$, C I $`\lambda \mathrm{\hspace{0.17em}1560}`$, respectively.
The higher resolution and signal-to-noise ratio of the UVES spectrum makes it possible to observe three new shallow components in C IV that are also present in H I Ly$`\alpha `$. As for the strongest feature, C IV $`\lambda \lambda \mathrm{\hspace{0.17em}1548},1550`$ and Si III $`\lambda \mathrm{\hspace{0.17em}1206}`$ are saturated and have been fitted following the velocity structure of Si IV $`\lambda \lambda \mathrm{\hspace{0.17em}1393},1402`$. The $`3\sigma `$ limit on the N V column density is $`N(`$N V$`)\stackrel{<}{}1.4\times 10^{12}`$ cm<sup>-2</sup>. The H I Ly$`\alpha `$, Ly$`\gamma `$, Ly$`ฯต`$ and Ly10 lines have been fitted with the same velocity structure as the C IV doublet for a total column density of $`N(`$H I$`)2\times 10^{16}`$ cm<sup>-2</sup> in agreement with the measurement by Prochaska and Burles (1999).
The low ionization lines C II $`\lambda \mathrm{\hspace{0.17em}1334}`$, Si II $`\lambda \lambda \mathrm{\hspace{0.17em}1190},1193`$, $`\lambda \mathrm{\hspace{0.17em}1260}`$,$`\lambda \mathrm{\hspace{0.17em}1304}`$, $`\lambda \mathrm{\hspace{0.17em}1526}`$, Mg II $`\lambda \lambda \mathrm{\hspace{0.17em}2803}`$, Al II $`\lambda \mathrm{\hspace{0.17em}1670}`$, and Al III $`\lambda \mathrm{\hspace{0.17em}1854}`$, have been detected and fitted with the same velocity profile. Mg II $`\lambda \lambda \mathrm{\hspace{0.17em}2796}`$ was not considered in the fit because its strongest component is affected by the presence of a telluric absorption line. We do not confirm the existence of the C I $`\lambda \lambda \mathrm{\hspace{0.17em}1560}`$ absorption tentatively identified by SAV98. $`N(`$Fe II$`)\stackrel{<}{}3.8\times 10^{11}`$ cm<sup>-2</sup> is the $`3\sigma `$ limit on the abundance of Fe II based on the absorption at $`\lambda \mathrm{\hspace{0.17em}2382}`$.
A more complete discussion on these latter two interesting metal systems will appear in a further paper (DโOdorico and Petitjean, 2000).
#### 5.1.21 System at $`z_{\mathrm{abs}}=2.07728`$
This weak C IV doublet was identified by SAV98.
We tentatively identified Si III $`\lambda \mathrm{\hspace{0.17em}1206}`$ and Fe II $`\lambda \mathrm{\hspace{0.17em}2382}`$ and a more reliable Si II $`\lambda \mathrm{\hspace{0.17em}1260}`$ absorption. The $`3\sigma `$ limit on the Si IV $`\lambda \mathrm{\hspace{0.17em}1393}`$ is: $`N(`$Si IV$`)\stackrel{<}{}3.2\times 10^{11}`$ cm<sup>-2</sup> and that on C II $`\lambda \mathrm{\hspace{0.17em}1334}`$, $`N(`$C II$`)\stackrel{<}{}1.3\times 10^{12}`$ cm<sup>-2</sup>. Unfortunately, the possible Mg II $`\lambda \lambda \mathrm{\hspace{0.17em}2796},2803`$ absorptions at this redshift fall in the gap between the two UVES red arm CCDs.
#### 5.1.22 System at $`z_{\mathrm{abs}}=2.11287`$
New, weak C IV doublet fitted with two components. The corresponding H I Ly$`\alpha `$ line shows a column density $`N(`$H I$`)8.1\times 10^{14}`$ cm<sup>-2</sup>.
### 5.2 Associated systems
The associated systems observed in the redshift range $`2.1982.206`$ have been investigated by Petitjean and Srianand (1999). They found that most of the lines in these systems show the signature of partial coverage and the covering factor varies from species to species. In general, absolute abundances are close to solar, with the exception of the \[N/C\] abundance ratio which is larger than solar. This result confirms the physical association of the absorbing gas with the AGN.
#### 5.2.1 System at $`z_{\mathrm{abs}}=2.19819`$
This associated system shows a strong C IV doublet at about $`2000`$ km s<sup>-1</sup> from the emission peak. OUT99 detected and fitted the N V doublet and the H I Ly$`\alpha `$ absorption line.
In the UVES spectrum we found a third, shallow component on the lower redshift side. C IV $`\lambda \lambda \mathrm{\hspace{0.17em}1548},1550`$, N V $`\lambda \lambda \mathrm{\hspace{0.17em}1238},1242`$, O VI $`\lambda \lambda \mathrm{\hspace{0.17em}1031},1037`$, H I Lya and H I Ly$`\beta `$ have been fitted together. The result of the fit is quite unsatisfactory ($`\chi ^2=9.5`$), due to the unusual ratio between the lines of the doublets. In particular, the O VI lines show a flat bottom - as if they were saturated - but a non zero residual flux, a clear signature of partial coverage. The $`3\sigma `$ limit on the Si IV $`\lambda \mathrm{\hspace{0.17em}1393}`$ is $`N(`$Si IV$`)\stackrel{<}{}5.2\times 10^{11}`$ cm<sup>-2</sup>.
#### 5.2.2 System at $`z_{\mathrm{abs}}=2.20106`$
This new, broad and shallow system shows the N V $`\lambda \lambda \mathrm{\hspace{0.17em}1238},1242`$ and C IV $`\lambda \mathrm{\hspace{0.17em}1548}`$ lines. C IV $`\lambda \mathrm{\hspace{0.17em}1550}`$ and O VI $`\lambda \lambda \mathrm{\hspace{0.17em}1031},1037`$ are blended in the forest. The $`3\sigma `$ limit on H I, from the absence of detectable H I Ly$`\alpha `$, is $`N(`$H I$`)\stackrel{<}{}3.6\times 10^{11}`$ cm<sup>-2</sup>.
#### 5.2.3 System at $`z_{\mathrm{abs}}=2.2064`$
A multi-component, broad C IV absorption in the redshift range between about $`1400`$ km s<sup>-1</sup> and $`1100`$ km s<sup>-1</sup> from the emission peak was observed by SAV98. OUT99 fitted the N V $`\lambda \lambda \mathrm{\hspace{0.17em}1238},1242`$ and the H I Ly$`\alpha `$ with three components, the fit is complicated by uncertainty in the precise shape of the Ly$`\alpha `$ & N V emission line continuum, also, the N V $`\lambda \mathrm{\hspace{0.17em}1242}`$ from the $`z_{\mathrm{abs}}=2.198`$ system lay on top of the N V $`\lambda \mathrm{\hspace{0.17em}1238}`$.
We performed a simultaneous fit of the C IV, N V, O VI doublets and the H I Ly$`\alpha `$ with five components. The $`3\sigma `$ limit on the Si IV column density is $`N(`$Si IV$`)\stackrel{<}{}3.3\times 10^{11}`$ cm<sup>-2</sup>.
We are greatly indebted to all the people involved in the conception, construction and commissioning of the UVES instrument, without whom this project would have been impossible. We are grateful to the members of the ESO Data Management Division Pascal Ballester, Sebastian Wolff and Andrea Modigliani for the development of the MIDAS-based UVES pipeline which was used to reduce these data. We thank J. Bergeron, S. DโOdorico, T.-S. Kim and S. Savaglio for enlightening discussions. This work has been conducted with partial support by the TMR programme Formation and Evolution of Galaxies set up by the European Community under the contract FMRX-CT96-0086.
|
warning/0006/nucl-th0006071.html
|
ar5iv
|
text
|
# Mott dissociation of D-mesons at the chiral phase transition and anomalous J/๐ suppression
## 1 Introduction
Recent results of the CERN NA50 collaboration on anomalous J/$`\psi `$ suppression in ultrarelativistic Pb-Pb collisions at 158 AGeV have renewed the quest for an explanation of the processes which may cause the rather sudden drop of the J/$`\psi `$ production cross section for transverse energies above $`E_T40`$ GeV in this experiment. An effect like this was predicted as a signal for quark gluon plasma formation due to screening of the $`c\overline{c}`$ interaction. Soon after that it became clear that for temperatures and densities just above the deconfinement transition the Mott effect for the J/$`\psi `$ does not occur and that a kinetic process as e.g. dissociation by meson or quark impact is required to dissolve the J/$`\psi `$ .
In this paper, we suggest that at the chiral/deconfinement phase transition the charmonium breakup reaction cross sections are critically enhanced since the open charm states of the dissociation processes become unbound (Mott effect) so that the reaction thresholds are effectively lowered. We present a model calculation for the particular process J/$`\psi +\pi D+\overline{D}^{}+h.c.`$ in a hot pion gas in order to demonstrate that the Mott dissociation of the D-mesons at the chiral phase transition leads to a threshold effect for the J/$`\psi `$ suppression ratio when applied to the case of Pb-Pb collisions at CERN within a modified Glauber model.
## 2 In-medium modification of charmonium break-up cross sections
The inverse lifetime of a charmonium state in a hot and dense many- particle system is given by the imaginary part of the selfenergy $`\tau ^1(p)=\mathrm{\Gamma }(p)=\mathrm{\Sigma }^>(p)\mathrm{\Sigma }^<(p)`$.
In the Born collision approximation for the dominant process in a hot pion gas, as shown in Fig. 1, we have
$`\mathrm{\Sigma }^{\stackrel{>}{<}}(p,\omega )`$ $`=`$ $`{\displaystyle _p^{}}{\displaystyle _{p_1}}{\displaystyle _{p_2}}(2\pi )^4\delta (p+p^{}p_1p_2)`$
$`\times `$ $`\left|\right|^2G_\pi ^{\stackrel{<}{>}}(p^{})G_{D_1}^{\stackrel{>}{<}}(p_1)G_{D_2}^{\stackrel{>}{<}}(p_2),`$
where the thermal Green functions $`G_i^>(p)=[1\pm f_i(p)]A_i(p)`$ and $`G_i^<(p)=f_i(p)A_i(p)`$ are defined by the spectral function $`A_i(p)`$ and the distribution function $`f_i(p)`$ of the state $`i`$; $`_p=\frac{d^4p}{(2\pi )^4}`$. In the low density approximation for the final states ($`f_{D_i}(p)0`$), one can safely neglect $`\mathrm{\Sigma }^<(p)`$ so that
$`\tau ^1(p)`$ $`=`$ $`{\displaystyle _p^{}}{\displaystyle _{p_2}}{\displaystyle _{p_2}}(2\pi )^4\delta (p+p^{}p_1p_2)`$
$`\times `$ $`\left|\right|^2f_\pi (p^{})A_\pi (p^{})A_{D_1}(p_1)A_{D_2}(p_2).`$
With the differential cross section
$$\frac{d\sigma }{dt}=\frac{1}{16\pi }\frac{\left|(s,t)\right|^2}{\lambda (s,M_\psi ^2,s^{})},$$
using $`s=(p+p^{})^2,t=(pp_1)^2,s^{}=p^2`$ and $`\sqrt{\lambda (s,M_\psi ^2,s^{})}=2v_{\mathrm{rel}}\sqrt{๐ฉ^2+M_\psi ^2}\sqrt{๐ฉ_{}^{}{}_{}{}^{2}+s^{}}`$ one can show that the $`J/\psi `$ relaxation time in a hot pion gas is given by
$`\tau ^1(p)={\displaystyle \frac{d^3๐ฉ^{}}{(2\pi )^3}๐s^{}}`$ $`f_\pi (๐ฉ^{},s^{})A_\pi (s^{})`$ (1)
$`\times `$ $`v_{\mathrm{rel}}\sigma ^{}(s),`$
where the pion spectral function decides whether the medium consists of resonant (off-shell) correlations or true $`q\overline{q}`$ bound states. The in-medium breakup cross section is given by
$$\sigma ^{}(s)=๐s_1๐s_2A_{D_1}(s_1)A_{D_2}(s_2)\sigma (s;s_1,s_2).$$
(2)
Note that there are two kinds of medium effects due to (i) the spectral functions of the final states and (ii) the explicit medium dependence of the matrix element $``$. In the following model calculation we will use the approximation $`\sigma (s;s_1,s_2)\sigma ^{\mathrm{vac}}(s;s_1,s_2)`$ justified by the locality of the transition matrix $``$ which makes it rather inert against medium influence.
## 3 Model calculation
The quark exchange processes in meson-meson scattering can be calculated within the diagrammatic approach of Barnes and Swanson which allows a generalization to finite temperatures in the thermodynamic Green function technique . This technique has been applied to the calculation of J/$`\psi `$ break-up cross sections by pion impact in , where also a fit formula has been given. The approach has been extended to excited charmonia states and consideration of rho-meson impact recently . The generic form of the resulting cross section (given a band of uncertainty) can be fit to the form
$$\sigma ^{\mathrm{vac}}(s;M_{D_1}^2,M_{D_2}^2)=\sigma _0\mathrm{ln}(s/s_0)\mathrm{exp}(s/\lambda ^2),$$
(3)
where $`s_0=(M_{D_1}+M_{D_2})^2`$ is the threshold for the process to occur, $`\sigma _0=7.510^9`$ mb and $`\lambda =0.9`$ GeV. An alternative fit is given in .
Recently, the charmonium dissociation processes have been calculated also in an effective Lagrangian approach , but this method introduces new phenomenological parameters and ignores the quark substructure. The development of a unifying approach on the basis of a relativistic confining quark model is in progress . Detailed numerical comparison cannot be made at present.
The major modification of the charmonium break-up process which we expect at finite temperatures in a hot meson gas comes from the Mott-effect for the light as well as the heavy mesons. At finite temperatures when the chiral symmetry in the light quark sector is restored, the continuum threshold for light-heavy quark pairs drops below the mass of the D-mesons so that they are no longer bound states constrained to their mass shell, but become rather broad resonant correlations in the continuum. This Mott-effect has been discussed within relativistic quark models for the light meson sector but can also be generalized to the case of heavy mesons . Applying a confining quark model we have obtained the critical temperatures $`T_D^{}^{\mathrm{Mott}}=110`$ MeV, $`T_D^{\mathrm{Mott}}=140`$ MeV.
In order to study the implications of the D-meson Mott effect for the charmonium breakup we adopt here a Breit-Wigner form for the spectral functions
$`A_i(s)`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle \frac{\mathrm{\Gamma }_i(T)M_i(T)}{(sM_i^2(T))^2+\mathrm{\Gamma }_i^2(T)M_i^2(T)}},`$ (4)
which in the limit of vanishing width $`\mathrm{\Gamma }_i(T)0`$ goes over into the delta function $`\delta (sM_i^2)`$ for a bound state in the channel $`i`$. The width of the D-mesons shall be modeled by a microscopic approach. For our exploratory calculation, we adopt here
$$\mathrm{\Gamma }_D(T)=c(TT_D^{\mathrm{Mott}})\mathrm{\Theta }(TT_D^{\mathrm{Mott}}),$$
(5)
where the coefficient $`c=2.67`$ is assumed to be universal for the pseudoscalar D-mesons and it is obtained from a fit to the pion width above the pion Mott temperature, see . For the D-meson mass we have $`M_D(T)=M_D+0.75\mathrm{\Gamma }_D(T)`$. The result for the in-medium J/$`\psi `$ break-up cross section (2) is shown in Fig. 2.
With $`M_D^{}=2.01`$ GeV and $`M_{\overline{D}}=1.87`$ GeV follows for the threshold $`s_0=15.05`$ GeV<sup>2</sup>. At a temperature $`T=140`$ MeV, where the D-meson can still be considered as a true bound state, the $`D^{}`$ meson has already entered the continuum and is a resonance with a half width of about 80 MeV. Due to the Mott effect for the open charm mesons, the charmonium dissociation processes become โsubthresholdโ ones and their cross sections which are peaked at threshold rise and spread to lower onset with cms energy. This is expected to enhance the rate for the charmonium dissociation processes in a hot meson gas.
## 4 J/$`\psi `$ dissociation in a hot pion gas
We calculate the inverse relaxation time for a J/$`\psi `$ at rest in a hot pion gas according to Eq. (1),
$`\tau ^1(T)`$ $`=`$ $`{\displaystyle \frac{d^3๐ฉ}{(2\pi )^3}f_\pi (๐ฉ)\frac{|๐ฉ|}{E_\pi (๐ฉ)}\sigma ^{}(s)}`$
$`=`$ $`<\sigma ^{}v_{\mathrm{rel}}>n_\pi (T),`$
where we have assumed on-shell pions with the dispersion relation $`E_\pi (๐ฉ)=\sqrt{๐ฉ^2+M_\pi ^2}`$ obeying the thermal Bose distribution function $`f_\pi (๐ฉ)=3[\mathrm{exp}[(E_\pi (๐ฉ)\mu )/T]1]^1`$, $`n_\pi (T)`$ being the pion density. The cms energy of the pion impact on a J/$`\psi `$ at rest is $`s(๐ฉ)=M_\pi ^2+M_\psi ^2+2M_\psi E_\pi (๐ฉ)`$. The result for the temperature dependence of the thermal averaged J/$`\psi `$ breakup cross section $`<\sigma ^{}v_{\mathrm{rel}}>`$ is shown in Fig. 3.
This quantity has to be compared to the nuclear absorption cross section for the J/$`\psi `$ of about 3 mb which has been extracted from charmonium suppression data in p-A collisions . It is remarkable that it is practically negligible below the D-meson Mott temperature $`T_D^{}^{\mathrm{Mott}}=110`$ MeV but comparable to the nuclear absorption cross section above the chiral/deconfinement temperature of $`T_{\mathrm{crit}}150`$ MeV. Therefore we expect the in-medium enhanced charmonium dissociation processes to be sufficiently effective to destroy the charmonium state on its way through the hot fireball of the heavy-ion collision and to provide an explanation of the observed anomalous J/$`\psi `$ suppression phenomenon .
## 5 Applications to Pb-Pb collisions at CERN
A more detailed comparison with the recent data from the NA50 collaboration requires a model for the heavy ion collision. In the following we will use a modified Glauber model which takes into accout the formation of secondaries (โsoft matterโ) during the collision and the propagation of the charmonium state through the debris of the collision. The suppression function is defined by
$`S(E_T)`$ $`=`$ $`S_N(E_T)\mathrm{exp}\left[{\displaystyle _{t_0}^{t_f}}๐t\tau ^1(n(t))\right]`$
$`=`$ $`S_N(E_T)\mathrm{exp}\left[{\displaystyle _{n_0(E_T)}^{n_f}}๐n<\sigma ^{}v_{\mathrm{rel}}>\right],`$
where we have assumed a a nuclear absorption systematics $`S_N(E_T)=18+36\mathrm{exp}(0.26\sqrt{E_T})`$ and longitudinal expansion according to $`n(t)=n_0(E_T)t_0/t`$. Within a Glauber model fit to the NA50 experiment we obtain the $`n_0(E_T)`$ as a parameter form depending on the impact parameter $`b`$, where $`E_T(b)/\mathrm{MeV}=130b/\mathrm{fm}`$ and $`n_0(b)/\mathrm{fm}^3=1.2\sqrt{1(b/10.8\mathrm{fm})^2}`$.
In Fig. 4 we show the result for our model calculation in comparison with the data for Pb-Pb collision experiments of the NA50 collaboration . The temperature (pion density) dependence of the J/$`\psi `$ breakup cross section which exhibits a critical enhancement at the D-meson Mott temperature, see Fig. 3 leads to a threshold behaviour in the $`E_T`$ dependence of the J/$`\psi `$ suppression ratio. This is in qualitative agreement with the data and the mechanism for the occurrence of thresholds in this quantity can be considered as a possible explanation for the observed deviation from the ordinary nuclear absorption systematics (anomalous J/$`\psi `$ suppression) in the NA50 experiment.
## 6 Summary and Outlook
In this contribution we have presented an approach to charmonium breakup in a hot and dense medium which is applicable in the vicinity of the chiral/deconfinement phase transition where mesonic bound states get dissolved in a Mott-type transition and should be described as resonant correlations in the quark plasma instead. This description can be achieved using the concept of the spectral function which can be obtained from relativistic quark models in a systematic way. The result of an exploratory calculation employing a temperature-dependent Breit-Wigner spectral function for open charm mesons presented in this paper has demonstrated that heavy-flavor dissociation processes are critically enhanced at the QCD phase transition and could lead in the charmonium sector to the phenomenon of anomalous J/$`\psi `$ suppression.
In subsequant work we will relax systematically approximations which have been made in the one presented here and improve inputs which have been used. In particular, we will investigate the off-shell behaviour of the charmonium breakup cross section in the vacuum (3) and calculate the spectral function (4) at finite temperature from a relativistic quark model. Dyson-Schwinger equations provide a nonperturbative, field-theoretical approach which has recently been applied also to heavy-meson observables and have proven successful for finite-temperature generalization . Further intermediate open charm states can be considered, the states in the dense environment should include rho mesons and nucleons besides of the pions which all can be treated as off-shell quark correlations at the QCD phase transition.
In future experiments at LHC the charm distribution in the created fireball may be not negligible so that the approximation $`f_{D_i}(p)0`$ has to be relaxed. In this case, not only the gain process ($`D\overline{D}`$ annihilation ) encoded in the $`\mathrm{\Sigma }^<`$ has to be included but at the same time the Bose enhancement factors in the $`G_i^>`$ functions have to be considered.
We want to emphasize that this approach to the kinetics of open charm and charmonium states in a dense medium provides a framework for a systematic investigation of processes relevant to QCD plasma diagnostics in heavy-ion collisions using heavy quark bound states.
## 7 Acknowledgements
This work has been supported by the Heisenberg-Landau program for scientific collaborations between Germany and the JINR Dubna and by the DFG Graduiertenkolleg โStark korrelierte Vielteilchensystemeโ at the University of Rostock. We thank T. Barnes, J. Hรผfner, M.A. Ivanov, V.N. Pervushin, C.D. Roberts, G. Rรถpke and P.C. Tandy for their discussions and stimulating interest in our work. D.B. acknowledges the hospitality of the INT Seattle where this work has been started, and of the ECT\* Trento where it has been completed.
|
warning/0006/math0006232.html
|
ar5iv
|
text
|
# 0. Introduction.
## 0. Introduction.
In this note I prove two results on the minimal generators of the defining ideals of closure of nilpotent conjugacy classes in the Lie algebra of $`n\times n`$ matrices. These results were motivated by questions of Pappas and Rapoport \[PR\] and they provide the answers to some of their conjectures.
In order to formulate the results let us recall the following notation.
Let $`K`$ be a field and let $`E`$ be a vector space of dimension $`n`$ over $`K`$. We denote by $`X=Hom_K(E,E)`$ the set of endomorphisms of $`E`$. Let $`\{e_1,\mathrm{},e_n\}`$ be a basis in $`E`$. We write the elements of $`X`$ in the form
$$\varphi (e_j)=\underset{i=1}{\overset{n}{}}\varphi _{i,j}e_i,$$
thus identifying $`X`$ with a set of $`n\times n`$ matrices. We also identify the space $`Hom_K(E,E)`$ with $`E^{}E`$. In this tensor product we identify $`\varphi `$ with
$$\varphi =\underset{i,j=1}{\overset{n}{}}\varphi _{i,j}e_j^{}e_i.$$
The general linear group $`GL(E)`$ acts by conjugation on $`X`$. For a partition $`\mu =(\mu _1,\mathrm{},\mu _s)`$ of $`n`$ let $`X_\mu `$ be the closure of the set $`O(\mu )`$ of nilpotent matrices with Jordan blocks of sizes $`\mu _1,\mathrm{},\mu _s`$. We denote by $`A=K[X]=Sym(EE^{})`$ the coordinate ring of $`X`$, and $`J(\mu )`$ denotes the defining ideal of $`X_\mu `$. The coordinate functions $`\mathrm{\Phi }_{i,j}`$ are the generators of $`A`$. They can be treated as independent indeterminates. We denote by
$$\mathrm{\Phi }=\left(\begin{array}{c}\mathrm{\Phi }_{i,j}\end{array}\right)_{1i,jn}$$
the generic $`n\times n`$ matrix.
Let us describe some important polynomials in $`A`$. Let $`I,J`$ be two subsets of $`[1,n]`$ of cardinality $`r`$. Then $`M(I,J)`$ denotes the determinant of the $`r\times r`$ submatrix of $`\mathrm{\Phi }`$ with rows from $`I`$ and columns from $`J`$.
We recall that the characteristic polynomial of $`\varphi `$ can be written
$$P(\varphi ,x)=\underset{i=0}{\overset{n}{}}T_ix^{ni},$$
where $`T_0=1`$ and $`T_i`$ is the sum of principal $`i\times i`$ minors of the generic matrix $`\mathrm{\Phi }`$,
$$T_i=\underset{I[1,n],card(I)=i}{}M(I,I).$$
The polynomials $`T_1,\mathrm{},T_n`$ are the generators of the ring of $`GL(E)`$-invariants in $`A`$.
We are interested in special nilpotent orbits. Let us fix a number $`e<n`$ and let $`n=re+f`$ be the division of $`n`$ by $`e`$ with the remainder $`f`$. To such $`n,e`$ we associate the partition $`\mu (n,e)=(e^r,f)`$. It is clear that the set $`X_{\mu (n,e)}`$ can be defined set theoretically by the conditions $`\varphi ^e=0`$. The subject of this note is the description of minimal generators of the ideal $`J(\mu (n,e))`$.
###### Theorem 1
Let $`K`$ be a field of characteristic zero. Then the ideal $`J(\mu (n,e))`$ is generated by the invariants $`T_1,\mathrm{},T_{e1}`$ and by the entries of the matrix $`\mathrm{\Phi }^e`$. These polynomials form a minimal set of generators of $`J(\mu (n,e))`$.
###### Theorem 2
Let $`K`$ be an arbitrary field and let $`e=2`$. The ideal $`J(\mu (n,2))`$ is generated by the invariants $`T_1,T_2,\mathrm{},T_n`$ and by the entries of the matrix $`\mathrm{\Phi }^2`$.
The proof of the first theorem is based on the geometric method used in \[W\]. The main result in \[W\] described a small but nonminimal set of generators of $`J(\mu )`$ for arbitrary $`\mu `$. Applying the method to the smaller class of partitions $`\mu (n,e)`$ yields Theorem 1.
The proof of Theorem 2 relies on the paper \[S\] of Strickland. Using that result one can easily show that the ideal $`J(n,2)`$ is generated by the polynomials described in Theorem 2 and by the minors of $`\mathrm{\Phi }`$ of rank $`[\frac{n}{2}]+1`$. It remains to show that the ideal generated by these minors is contained in the ideal generated by the polynomials described in Theorem 2. This is done by explicit calculation, using some information about two-rowed Schur functors in characteristic $`p>0`$.
Throughout the paper we denote by $`S_\mu E`$ the Schur module corresponding to the highest weight $`\mu `$ for the group $`GL(n)`$. This means that $`\mu =(\mu _1,\mathrm{},\mu _n)`$ where $`\mu _i๐`$, $`\mu _1\mathrm{}\mu _n`$.
## 1. The proof of Theorem 1.
Let $`E,K`$ be as above. For a given number $`p`$, $`1pn`$ we consider the subrepresentation of $`A`$ which is the span of $`p\times p`$ minors of the generic matrix $`\mathrm{\Phi }`$. This vector space can be identified with $`^pE^pE^{}`$. For arbitrary $`i`$, $`0ip`$ we define the $`GL(E)`$-equivariant map $`\theta (i,p)`$ to be the composition
$$\stackrel{i}{}E\stackrel{i}{}E^{}\stackrel{i}{}E\stackrel{i}{}E^{}K\stackrel{1tr^{\left(pi\right)}}{}\stackrel{i}{}E\stackrel{i}{}E^{}\stackrel{pi}{}E\stackrel{pi}{}E^{}\stackrel{p}{}E\stackrel{p}{}E^{}$$
where the second map is tensoring with the $`(pi)`$โst divided power of the trace element $`tr=_{i=1}^ne_ie_i^{}EE^{}`$ and the third map is a product of two exterior multiplications.
We define $`V_{i,p}`$ to be the image of this map. When $`K`$ is a field of characteristic zero it is well known that the space $`^pE^pE^{}`$ has the following $`GL(E)`$-equivariant decomposition into irreducible representations of $`GL(E)`$,
$$\stackrel{p}{}E\stackrel{p}{}E^{}=_{j=0}^{min(p,np)}S_{1^j,0^{n2j},(1)^j}E,$$
where $`S_{1^j,0^{n2j},(1)^j}E`$ denotes the irreducible representation (Schur functor) with the highest weight $`(1^j,0^{n2j},(1)^j)`$. We denote the summand $`S_{1^j,0^{n2j},(1)^j}E`$ occurring in $`^pE^pE^{}`$ by $`U_{j,p}`$.
Notice that $`U_{0,p}=V_{0,p}`$ can be identified with the one dimensional subspace generated by the invariant $`T_p`$.
###### Lemma 1
Let $`K`$ be a field of characteristic zero. We have $`V_{i,p}=_{j=0}^{min(i,np)}U_{j,p}`$.
Proof. This is an easy calculation after one observes that the highest weight vector in $`^pE^pE^{}`$ of weight $`(1^i,0^{2ni},(1)^i)`$ is
$$w(i,p)=\underset{J[1,n],|J|=pi}{}(e_1\mathrm{}e_ie_J)(e_J^{}e_{n+1i}^{}\mathrm{}e_n^{}),$$
where for $`J=\{j_1,\mathrm{},j_t\}`$ we write $`e_J=e_{j_1}\mathrm{}e_{j_t}`$, $`e_J^{}=e_{j_1}^{}\mathrm{}e_{j_t}^{}`$.$``$
###### Remark
a) Notice that $`U_{0,p}=V_{0,p}`$ can be identified with the one dimensional subspace generated by the invariant $`T_p`$. The representation $`V_{1,p}`$ is the subspace generated by the entries of the $`p`$-th power $`\mathrm{\Phi }^p`$ of the generic matrix.
b) We have a commutative diagram
$$\begin{array}{ccc}^{i1}E^{i1}E^{}& \stackrel{1T_1}{}& ^iE^iE^{}\\ \\ \left(pi+1\right)\theta (i1,p)& & \theta (i,p)\\ \\ & ^pE^pE^{}& \end{array}$$
which proves that $`V_{i1,p}V_{i,p}`$.
Before we start with the proof we need a couple of elementary lemmas.
###### Lemma 2
Let $`i1`$. Then the vector space $`V_{i,p+1}`$ is contained in the ideal generated by $`V_{i,p}`$.
Proof. Let us take a typical element of $`V_{i,p+1}`$. It can be written as
$$\underset{j_1,\mathrm{},j_{p+1i}}{}M(u_1,\mathrm{}u_i,j_1,\mathrm{},j_{p+1i};v_1,\mathrm{},v_i,j_1,\mathrm{},j_{p+1i}).$$
Here we are summing over all choices of $`j_1,\mathrm{},j_{p+1i}`$ regardless of whether the summands are nonzero. Let us take the Laplace expansion of every term in the above sum, with respect to the row $`u_1`$. This allows us to express our generator in the following form
$$\underset{j_1,\mathrm{},j_{p+1i}}{}(\underset{k=1}{\overset{i}{}}(1)^{k+1}\mathrm{\Phi }_{u_1,v_k}M(u_2,\mathrm{}u_i,j_1,\mathrm{},j_{p+1i};v_1,\mathrm{},\widehat{v}_k\mathrm{},v_i,j_1,\mathrm{},j_{p+1i})+$$
$$+\underset{l=1}{\overset{p+1i}{}}(1)^{i+l+1}\mathrm{\Phi }_{u_1,j_l}M(u_2,\mathrm{}u_i,j_1,\mathrm{},j_{p+1i};v_1,\mathrm{},v_i,j_1,\mathrm{},\widehat{j}_l,\mathrm{},j_{p+1i})).$$
Writing the first part of the sum as a linear combination of $`\mathrm{\Phi }_{u_1,v_k}`$ we see that all coefficients are in $`V_{i1,p}V_{i,p}`$. Similarly, writing the second part of the sum as a linear combination of $`\mathrm{\Phi }_{u_1,j_l}`$ we see that each coefficient is in $`V_{i,p}`$.$``$ We also recall the geometric method for calculating syzygies. We consider the Springer type desingularization $`Y_\mu `$ of $`X_\mu `$. Let $`\mu ^{}=(\mu _1^{},\mathrm{},\mu _t^{})`$ be the partition conjugate to $`\mu `$. Let $`G/P_\mu ^{}`$ be the flag variety of partial flags $`(R_1\mathrm{}R_{t1}R_t=E)`$ with $`dimR_i=\mu _1^{}+\mathrm{}+\mu _i^{}`$, and
$$Y_\mu =\{(\varphi ,R_1,\mathrm{},R_t)X\times G/P_\mu ^{}|\varphi (R_i)R_{i1}\mathrm{for}i=1,\mathrm{},t\}.$$
We denote by $`_i`$ the tautological vector bundle on $`G/P_\mu ^{}`$ of dimension $`\mu _1^{}+\mathrm{}+\mu _i^{}`$, corresponding to the subspace $`R_i`$. We denote by $`๐ฌ_i`$ the tautological factorbundle $`(E\times G/P_\mu ^{})/_i`$.
The variety $`Y_\mu `$ is the total space of the cotangent bundle on $`G/P_\mu ^{}`$. Let $`๐ฏ_\mu `$ denote the tangent bundle on $`G/P_\mu `$. Then we have an exact sequence of vector bundles on $`G/P_\mu ^{}`$
$$0๐ฎ_\mu EE^{}\times G/P_\mu ^{}๐ฏ_\mu 0.$$
The bundle $`๐ฎ_\mu `$ is โthe staircase bundleโ. In terms of tautological bundles $`๐ฎ_\mu `$ can be described (as a subbundle of $`(EE^{})\times G/P_\mu ^{}`$) as
$$๐ฎ_\mu =_1๐ฌ_0^{}+_2๐ฌ_1^{}+\mathrm{}+_t๐ฌ_{t1}^{}$$
(the sum is not direct), with the convention $`_t=E\times G/P_\mu ^{}`$, $`๐ฌ_0=E^{}\times G/P_\mu ^{}`$.
Recall that the basic theorem of the geometric method (comp. \[W\], \[F-W\]) implies the following
###### Theorem 3
The varieties $`X_\mu `$ are normal, with rational singularities. Moreover, a minimal free resolution of the coordinate ring $`K[X_\mu ]`$ as an $`A`$-module has the terms
$$\mathrm{}F_i^\mu F_{i1}^\mu \mathrm{}F_1^\mu F_0^\mu 0$$
where
$$F_i^\mu =_{j0}H^j(G/P_\mu ,\stackrel{i+j}{}๐ฎ_\mu )_KA(ij).$$
where $`A(d)`$ denotes the one dimensional homogeneous free $`A`$-module with a generator in degree $`d`$.
Let $`\widehat{\mu }`$ be the partition $`\widehat{\mu }=(max(\mu _11,0),max(\mu _21,0),\mathrm{},max(\mu _t1,0))`$. Then $`\widehat{\mu }^{}=(\mu _2^{},\mathrm{},\mu _t^{})`$. We see that we have an exact sequence
$$0E^{}_1๐ฎ_\mu ๐ฎ_{\widehat{\mu }}(๐ฌ_1,๐ฌ_1^{})0$$
$`()`$
where $`๐ฎ_{\widehat{\mu }}(๐ฌ_1,๐ฌ_1^{})`$ is the bundle $`๐ฎ_{\widehat{\mu }}`$ constructed in a relative situation with $`๐ฌ_1`$ replacing $`E`$.
Using the exact sequence $`()`$ we can estimate the terms of the complex $`F_{}^\mu `$ inductively. We consider the minimal resolution $`F^{\widehat{\mu }}(๐ฌ_1,๐ฌ_1^{})_{}`$ of the nilpotent orbit closure for the partition $`\widehat{\mu }`$ in the relative situation (i.e. $`๐ฌ_1`$ replaces $`E`$ and we treat the resolution as the sequence of $`Sym(๐ฌ_1๐ฌ_1^{})`$-modules). For each term $`S_\lambda ๐ฌ_1`$ occurring in that resolution we can describe the push down $`K_\lambda (E,E^{})`$ of the complex $`S_\lambda ๐ฌ_1^{}(E^{}_1)`$. The terms of the complexes $`K_\lambda (E,E^{})`$ for all representations $`S_\lambda ๐ฌ_1`$ occurring in $`F^{\widehat{\mu }}(๐ฌ_1,๐ฌ_1^{})_{}`$ give an upper bound on the terms from $`F_{}^\mu `$.
The following facts are proved in \[W\], section 3.
###### Theorem 4
Let $`\mu `$, $`\widehat{\mu }`$ be as above. Let $`S_\lambda ๐ฌ_1`$ be the term in $`F^{\widehat{\mu }}(๐ฌ_1,๐ฌ_1^{})_i`$. Then the terms coming from $`K_\lambda (E,E^{})`$ appear in homological degrees $`i`$, i.e. they can give the contribution only to the terms $`F_j^\mu `$ with $`ji`$.
In particular, as stated in Corollary (3.15) of \[W\], the term $`F_0^\mu `$ is trivial, consisting of a copy of $`A`$ in homogeneous degree $`0`$. The terms in $`F_1^\mu `$ are, by Theorem (3.12) of \[W\], of two types:
1. The term $`F^{\widehat{\mu }}(๐ฌ_1,๐ฌ_1^{})_0`$ is trivial, so the corresponding complex $`K_{(0)}`$ is just a pushdown of $`^{}(E^{}_1)`$. Its first term gives a possible contribution to $`F_1^\mu `$ which is the vanishing of $`n\mu _1+1`$ minors of the matrix $`\mathrm{\Phi }`$,
2. Each term $`S_\lambda ๐ฌ_1`$ from $`F^{\widehat{\mu }}(๐ฌ_1,๐ฌ_1^{})_1`$ gives a possible contribution to $`F_1^\mu `$ which is the zeroโth term of the corresponding complex $`K_\lambda `$. Now let us use the above inductive procedure for the family of partitions $`\mu (n,e)=(e^r,f)`$. The partition $`\widehat{\mu }(n,e)=((e1)^r,max(f1,0))`$ is just $`\mu (nr1,e1)`$ for $`f>0`$ or $`\mu (nr,e1)`$ for $`f=0`$. These partitions are in the same family, so we can assume by induction that Theorem 1 is true for $`\widehat{\mu }`$.
The inductive assumption means that the first term of $`F_1^{\widehat{\mu }(n,e)}`$ consists of the trivial representations in homogeneous degrees $`1,\mathrm{},e1`$ and the representation $`S_{1,0,\mathrm{},0,1}๐ฌ_1`$ in homogeneous degree $`e1`$. This means that the terms in $`F_1^\mu `$ have trivial representations in homogeneous degrees $`1,\mathrm{},e1`$, the representation $`E^{}E`$ in degree $`e`$ (comp. Lemma (3.11) in \[W\]) and the terms coming from $`n\mu (n,e)_1+1=nr`$ size minors of the matrix $`\mathrm{\Phi }`$. The generators in degrees $`e`$ have to consist of the invariants $`T_1,\mathrm{},T_{e1}`$ and the vanishing of entries of $`\mathrm{\Phi }^e`$, because these terms have to occur in $`F_1^\mu `$ and they match the terms we described. Thus the induction implies that the minimal generators of $`J(\mu (n,e))`$ are those listed in Theorem 1 plus possibly some linear combinations of $`(nr)\times (nr)`$ minors of $`\mathrm{\Phi }`$.
However we also recall what the Theorem (4.6) from \[W\] says about the ideal $`J(n,r)`$.
###### Theorem 5
The ideal $`J(\mu (n,e))`$ is generated (nonminimally) by the invariants $`U_{0,p}`$ ($`1pn`$) and by the representations $`U_{i,iei+1}`$ (for $`1ir`$).
The only possible generator in degree $`nr`$ on the above list is the invariant $`U_{0,nr}`$. However this cannot be a minimal generator because $`U_{0,nr}`$ is contained in $`V_{1,nr}`$ which is in the ideal generated by $`V_{1,e}`$ according to Lemma 2. This concludes the proof of Theorem 1.
## 2. The proof of Theorem 2.
In this section we work over a field $`K`$ of arbitrary characteristic. We are interested in the generators of the ideal $`J(\mu (n,2))`$. Let us recall that Strickland in \[S\] exhibited a standard basis in the coordinate ring $`A/J(\mu (n,2))`$ of the orbit closure $`X_{\mu (n,2)}`$. Denote by $`M(a_1,\mathrm{},a_r;b_1,\mathrm{},b_r)`$ the $`r\times r`$ minor of $`\mathrm{\Phi }`$ corresponding to rows $`a_1,\mathrm{},a_r`$ and columns $`b_1,\mathrm{},b_r`$. One notices easily that the only relations used in \[S\] to prove that the standard monomials span the ring $`A/J(\mu (n,2))`$ are the relations
$$\underset{1i_1<\mathrm{}<i_rn}{}M(a_1,\mathrm{},a_{pr},i_1,\mathrm{},i_r;b_1,\mathrm{},b_{pr},i_1,\mathrm{},i_r)$$
$`Rel(r,p)`$
for $`1pn,1rp`$ and $`([\frac{n}{2}]+1)\times ([\frac{n}{2}]+1)`$ minors of $`\mathrm{\Phi }`$. Therefore Theorem 2 is a consequence of the following lemmas.
###### Lemma 3
The polynomials of type $`Rel(r,p)`$ for $`1pn,1rp`$ are in the ideal generated by the invariants $`T_1,T_2,\mathrm{},T_n`$ and by the entries of $`\mathrm{\Phi }^2`$.
Proof. Let us denote by $`J`$ the ideal generated by the invariants $`T_1,\mathrm{},T_n`$ and by the entries of $`\mathrm{\Phi }^2`$. We use induction on $`p`$. For $`p=1,2`$ we see that the equations of type $`Rel(r,p)`$ describe exactly $`T_1`$ and the entries of $`\mathrm{\Phi }^2`$. Let us assume that the statement is true for $`p`$. Let us first assume that $`r>1`$. If $`pr=0`$ we are dealing with the invariant $`T_p`$, so it is in $`J`$. If $`pr>0`$ we see that after taking the Laplace expansion of $`Rel(r,p+1)`$ with respect to the row $`a_1`$ we get a linear combination of terms of type $`Rel(r1,p)`$ so we are done by induction. For $`r=1`$ we note that
$$\underset{1in}{}M(a_1,\mathrm{},a_p,i;b_1,\mathrm{},b_p,i)=$$
$$=\underset{1in}{}\underset{1up}{}(1)^{u+1}M(a_1;b_u)M(a_2,\mathrm{},a_p,i;b_1,\mathrm{},\widehat{b}_u,\mathrm{},b_p,i)+$$
$$+(1)^p\underset{1in}{}M(a_1;i)M(a_2,\mathrm{},a_p,i;b_1,\mathrm{},b_p).$$
The first summand on the right hand side is in $`J`$ by induction. It remains to deal with the second summand which we will denote by $`F`$. We take the Laplace expansion with the respect to the row $`i`$ in each term in that summand. We obtain
$$F=\underset{1in}{}\underset{1vp}{}(1)^{v+1}M(a_1;i)M(i;b_v)M(a_2,\mathrm{},a_p;b_1,\mathrm{},\widehat{b}_v,\mathrm{},b_p).$$
However
$$\underset{1in}{}M(a_1;i)M(i;b_v)=\underset{1in}{}M(i;i)M(a_1;b_v)+\underset{1in}{}M(a_1,i;b_v,i).$$
Since the second summand above is in $`J`$, we see that in $`A/J`$ we have
$$F=\underset{1in}{}M(i,i)\underset{1vp}{}(1)^{v+1}M(a_1,b_v)M(a_2,\mathrm{},a_p;b_1,\mathrm{},\widehat{b}_v,\mathrm{},b_p)=$$
$$=T_1M(a_1,\mathrm{},a_p;b_1,\mathrm{},b_p)$$
so $`FJ`$ and we are done.
###### Lemma 4
The $`([\frac{n}{2}]+1)\times ([\frac{n}{2}]+1)`$ minors of $`\mathrm{\Phi }`$ are contained in the ideal spanned by the invariants $`T_1,T_2,\mathrm{},T_n`$ and by the entries of the matrix $`\mathrm{\Phi }^2`$.
Proof. We employ the following notation. We write $`m=[\frac{n}{2}]+1`$. We identify the set of $`m\times m`$ minors with $`^mE^mE^{}=^mE^{nm}E`$. This is only an $`SL(E)`$-isomorphism but it is sufficient for our purposes. The relation $`Rel(r,m)`$ translates under this identification to the image of the following map.
$$\psi (r,m):\stackrel{mr}{}E\stackrel{nm+r}{}E\stackrel{1\mathrm{\Delta }}{}\stackrel{mr}{}E\stackrel{r}{}E\stackrel{nm}{}E\stackrel{m1}{}\stackrel{m}{}E\stackrel{nm}{}E$$
Therefore we need to prove
###### Lemma 5
The vector space $`^mE^{nm}E`$ is spanned by the images of the maps $`\psi (r,m)`$ for $`1rm`$.
Proof. We use induction on $`n`$. We prove the spanning weight by weight. Notice that the only weights in $`^mE^{nm}E`$ are (up to permutation of the basis) the weights $`(2^j,1^{n2j},0^j)`$. Notice that the problem of spanning in the weight $`(2^j,1^{n2j},0^j)`$ for a given $`n`$ is equivalent to the problem of spanning in the weight $`(1^{n2j})`$ for the dimension $`n2j`$. So by induction on $`n`$ we can assume that all vectors of weights $`(1^n)`$ are in the subspace generated by the images of $`\psi (r,m)`$. But the subrepresentation of $`^mE^{nm}E`$ generated by all vectors of weights $`(1^n)`$ is isomorphic to the kernel of the exterior multiplication
$$\stackrel{m}{}E\stackrel{nm}{}E\stackrel{n}{}E.$$
Therefore it is enough to show that the images of $`\psi (r,m)`$ cannot be all contained in this kernel. However one sees easily that the map $`\psi (r,m)`$ composed with the exterior multiplication is the exterior multiplication
$$\stackrel{mr}{}E\stackrel{nm+r}{}E\stackrel{n}{}E$$
multiplied by the binomial coefficient $`\left(\genfrac{}{}{0pt}{}{nm+r}{r}\right)`$. So Theorem 2 follows from the last elementary lemma.
###### Lemma 6
Let $`n`$ be an integer, $`m=[\frac{n}{2}]+1`$. Then
$$GCD_{1rm}\left(\genfrac{}{}{0pt}{}{nm+r}{r}\right)=1.$$
Proof. Let $`n=2t`$ be even. Then $`m=t+1`$. We are interested in
$$GCD_{1rt+1}\left(\genfrac{}{}{0pt}{}{t1+r}{r}\right).$$
Using the relations $`\left(\genfrac{}{}{0pt}{}{k}{r}\right)=\left(\genfrac{}{}{0pt}{}{k1}{r}\right)+\left(\genfrac{}{}{0pt}{}{k1}{r1}\right)`$ we get
$$GCD_{2rt+1}\left(\genfrac{}{}{0pt}{}{t1+r}{r}\right)=GCD_{2rt+1}\left(\genfrac{}{}{0pt}{}{t+1}{r}\right)$$
which obviously equals to $`1`$.
If $`n=2t+1`$ is odd, then $`m=t+1`$ and we are interested in
$$GCD_{1rt+1}\left(\genfrac{}{}{0pt}{}{t+r}{r}\right).$$
Again using the same method we find that
$$GCD_{1rt+1}\left(\genfrac{}{}{0pt}{}{t+r}{r}\right)=GCD_{1rt+1}\left(\genfrac{}{}{0pt}{}{t+1}{r}\right)=1.$$
## References.
\[FW\] T. Fukui, J. Weyman, Cohen-Macaulay properties of Thom-Boardman strata. I. Morinโs ideal. Proc. London Math. Soc. (3) 80 (2000), no. 2, 257-303,
\[J\] G. James, Representations of the Symmetric Group, Springer LN, 628,
\[PR\] G. Pappas, M. Rapoport, Local models in the ramified case. I. The EL-case, preprint, 2000,
\[S\] E. Strickland, On the variety of projectors, J. Algebra, 106(1987), 135-147,
\[W\] J. Weyman, The equations of conjugacy classes of nilpotent matrices , Invent. Math. 98, 229-245 (1989),
Department of Mathematics
Northeastern University
Boston, MA, 02115
USA
e-mail:weyman@research.neu.edu
|
warning/0006/nucl-ex0006003.html
|
ar5iv
|
text
|
# The Halo of 14Be
## Abstract
The two-neutron halo nucleus <sup>14</sup>Be has been investigated in a kinematically complete measurement of the fragments (<sup>12</sup>Be and neutrons) produced in dissociation at 35 MeV/nucleon on C and Pb targets. Two-neutron removal cross-sections, neutron angular distributions and invariant mass spectra characteristic of a halo were observed and the electromagnetic (EMD) contributions deduced. Comparison with three-body model predictions indicate that the halo wavefunction contains a large $`\nu (2s_{1/2})^2`$ admixture. The EMD invariant mass spectrum exhibited a relatively narrow structure near threshold (E<sub>decay</sub>=1.8$`\pm `$0.1 MeV, $`\mathrm{\Gamma }`$ = 0.8$`\pm `$0.4 MeV) consistent with a soft-dipole excitation.
PACS number(s): 27.20.+n6, 25.60.Dz, 25.60.Ge, 24.30.Gd
The size and distribution of matter in the nucleus have long played a central role in nuclear physics. Indeed, such gross properties reflect the combined effects of many fundamental aspects of nuclei. For the stable nuclei, measurements employing conventional probes, such as high energy electron and hadron scattering, have shown that the neutron and proton distributions exhibit essentially identical radii . In contrast, for some light nuclei far from stability, which combine a large neutron excess with very weak binding, large differences have been found. Such โhaloโ systems are well described by a core, resembling a normal nucleus, surrounded by an extended valence neutron density distribution .
In general terms the halo may be regarded as a threshold phenomenon whereby the loosely bound valence neutrons tunnel with significant probability into the classically forbidden region outside the core potential. Within a simple quasideuteron description, the extent of the halo is governed by the separation energy and reduced mass of the system . Under more realistic considerations the development of the halo is also influenced by the centrifugal barrier . In the cases of <sup>6,8</sup>He, <sup>11</sup>Be and <sup>11</sup>Li, which have been investigated experimentally in considerable detail, the valence neutrons occupy the $`2s_{1/2}`$ and/or $`1p_{3/2,1/2}`$ single-particle orbitals. In <sup>14</sup>Be the configuration of the halo neutrons would, in a naรฏve shell model prescription, be $`\nu (d_{5/2})^2`$. Sophisticated models suggest, however, that a $`\nu (s_{1/2})^2`$ admixture is also present . Unfortunately, a paucity of experimental data has precluded the elucidation of the structure of <sup>14</sup>Be beyond the matter radius . Compared to the other halo systems, the comparatively strong binding of the valence neutrons in <sup>14</sup>Be (S<sub>2n</sub>=1.34$`\pm `$0.11MeV ) combined with the $`\nu (d_{5/2})^2`$ component may provide a new window on continuum excitations, including the long sought-after Soft-Dipole Resonance (SDR) .
The goal of the present study was thus to explore the halo structure and continuum excitations of the two-neutron halo nucleus <sup>14</sup>Be. The tool chosen was a kinematically complete measurement of the fragments (<sup>12</sup>Be and two neutrons) from the dissociation of an intermediate energy beam of <sup>14</sup>Be on C and Pb targets. Such a measurement allowed the two-neutron removal cross sections, neutron angular distributions and invariant mass spectra to be extracted (the results of an analysis of the neutron-neutron correlations have been presented elsewhere ). The use of C and Pb targets permitted the electromagnetic component of the dissociation (EMD) to be deduced.
The <sup>14</sup>Be beam ($``$130 pps) was prepared using the LISE3 spectrometer and a 63 MeV/nucleon <sup>18</sup>O primary beam bombarding a thick Be production target. The mean energy of the beam at the mid-point of the secondary breakup targets was 35 MeV/nucleon. The energy spread in the beam was 10% and was compensated for by a time-of-flight (TOF) measurement over a 24 m flight-path between a parallel-plate avalanche counter (PPAC) located at the first focus of the spectrometer and the beam identification Si-detector. The beam particles were tracked onto the breakup targets (C 275 mg/cm<sup>2</sup>, Pb 570 mg/cm<sup>2</sup>) using two position sensitive PPACโs (resolution FWHM $``$ 1-2 mm). Owing to the mixed nature of the secondary beam (50 % <sup>14</sup>Be) the incoming ions were identified on a particle-by-particle basis using the TOF information combined with the energy loss derived from a Si-detector (300 $`\mu `$m) located just upstream of the target. The charged fragments from breakup were identified using a large area (5$`\times `$5 cm<sup>2</sup>) position sensitive (FWHM $``$ 0.5mm) Si-CsI telescope (Si 500 $`\mu `$m, CsI 2.5 cm) centred at zero degrees and located 11.4 cm downstream of the target. The energy response of the telescope (FWHM = 1.5%) was calibrated using various mixed secondary beams containing <sup>12</sup>Be with energies straddling that expected for <sup>12</sup>Be fragments arising from the dissociation of <sup>14</sup>Be. In order to account for events arising from reactions in the telescope, data was also acquired without a reaction target with the beam energy reduced by the amount corresponding to the energy loss in the C and Pb targets.
The neutrons emitted at forward angles were detected using the 99 elements of the DรฉMoN array . The array covered angles between +13 and -40 in the horizontal plane and $`\pm `$14 in the vertical with the modules arranged in a staggered configuration at distances between 2.5 and 6.5 m from the target . Such a geometry provided for a relatively high two-neutron detection efficiency (1.5%) whilst reducing the rate of cross-talk โ both intrinsically and via the use of an off-line rejection algorithm โ to negligible levels . A threshold of 15 MeV on the neutron energy was applied in the off-line analysis to eliminate contamination from the small number of evaporation neutrons arising from the target.
The results obtained for the two-neutron removal cross sections, $`\sigma _{2n}`$ (<sup>12</sup>Be identified in the telescope), the single-neutron angular distributions, $`d\sigma /d\mathrm{\Omega }`$ (<sup>12</sup>Be and neutron), and the associated angle integrated (0-40) cross sections, $`\sigma _n`$, are displayed in table I and figure 1a. In addition, the average neutron multiplicities have been derived ($`\overline{m}_n=\sigma _n/\sigma _{2n}`$) and are also listed. The single-neutron angular distributions are well characterised by a Lorentzian lineshape and the corresponding momentum width parameters, $`\mathrm{\Gamma }_n`$, have been tabulated. The large neutron removal cross sections and relatively narrow neutron distributions, while not as pronounced as for <sup>11</sup>Li , clearly indicate the halo character of <sup>14</sup>Be. The present results improve considerably on the earlier measurements of Riisager et al. which suffered from poor statistics (no angular distribution could be constructed for a heavy target) and were restricted to a limited angular range.
The multiplicities obtained for the two targets are instructive in terms of the reaction mechanisms leading to dissociation . For a light target, unless the halo neutrons are highly spatially correlated, the reaction is expected to proceed via single-neutron removal (absorption or diffraction) followed by the in-flight decay of <sup>13</sup>Be. As approximately equal contributions are expected for absorption and diffraction the average neutron multiplicity should be 1.5, in accordance with that measured here (table I). This scenario is also supported by the single-neutron angular distribution for the C target which is well reproduced assuming passage via a low-lying resonance in <sup>13</sup>Be . In the case of a heavy target, nuclear and Coulomb dissociation are present. Given that Coulomb dissociation should be associated with a multiplicity of 2, the average multiplicity for dissociation on Pb should be between 1.5 and 2, as observed.
The enhanced cross section for dissociation on the Pb target is indicative of a large EMD contribution. Assuming that the nuclearโCoulomb interference is small, the C target data (which arises essentially from nuclear induced reactions) may be scaled to estimate the nuclear contribution to breakup on Pb . Assuming a root-mean-square radius of 3.2 fm for <sup>14</sup>Be , $`\sigma _{2n}^{nucl}(Pb)`$ = 0.85$`\pm `$0.07 b and, consequently, $`\sigma _{2n}^{EMD}(Pb)`$ = 1.45$`\pm `$0.40 b. The latter can be compared to the value of 0.47$`\pm `$0.15 b measured at 800 MeV/nucleon . Importantly, for halo nuclei, the EMD cross section is dominated by the E1 component . An enhancement with decreasing beam energy is thus expected, owing to the large amount of dipole strength near threshold (see below) coupled with the weighting of the virtual photon spectrum to low photon energies .
Assuming that the neutron angular distribution arising from nuclear dissociation on Pb is identical to that measured for the C target, the single-neutron angular distribution for EMD has been constructed (figure 1b) and the corresponding integrated cross section and average multiplicity derived (table I). Interestingly, the angular distribution remains narrow and forward peaked whilst the multiplicity is consistent with the value of 2 expected for EMD, confirming the validity of the methods used to estimate the contribution arising from nuclear breakup.
The invariant mass spectra, reconstructed from the measured momenta of the beam and fragments (<sup>12</sup>Be and two neutrons) from breakup, are displayed in figure 2a and b for the C and Pb targets. The EMD spectrum (figure 2c) has been deduced, as described above, following subtraction of the estimated nuclear contribution to reactions on Pb. As for the spectra obtained with the C and Pb targets, the EMD spectrum exhibits enhanced strength around 2 MeV decay energy ($`E_{decay}`$). Given the complex nature of the response function of the present setup, a detailed Monte Carlo simulation, including the influence of all nonactive materials, was developed based on the GEANT package . The results shown in figure 2 were obtained following the descriptions for dissociation on C and Pb outlined earlier. In the case of the nuclear induced reactions a single low-lying state in <sup>13</sup>Be ($`E_0=0.5`$ MeV, $`\mathrm{\Gamma }_0=0.5\pm 0.4`$ MeV) was assumed to be populated following the diffraction of one of the halo neutrons . The EMD was simulated under the assumption that the energy sharing between the <sup>12</sup>Be and the two neutrons was governed by 3-body phase space. As shown in figure 2c, the observed EMD decay energy spectrum could be reproduced using a Breit-Wigner lineshape with a resonance decay energy of $`E_0=1.8\pm 0.1`$ MeV and width $`\mathrm{\Gamma }_0=0.8\pm 0.4`$ MeV. Furthermore, the corresponding simulations of the single-neutron angular distributions were in good agreement with those observed for reactions on C and Pb, as well as that deduced for EMD .
As noted above, the EMD of halo nuclei is essentially E1 in character. An analytical estimate for the E1 strength for two-neutron halo nuclei has been derived in a simple 3-body model based on Yukawa wavefunctions , whereby the maximum occurs for $`E_{decay}=6/5S_{eff}`$, where $`S_{eff}1.5S_{2n}`$. Whilst agreeing well with the available results for <sup>11</sup>Li, a maximum is predicted for <sup>14</sup>Be at $`E_{decay}2.4`$ MeV, somewhat above that observed here.
Thompson and Zhukov have examined <sup>14</sup>Be within the framework of a more realistic 3-body model in which the <sup>12</sup>Be core is treated as inert and a number of trial wavefunctions developed. Based on the binding energy and matter radius of <sup>14</sup>Be, together with the known d-wave resonance at 2.01 MeV in <sup>13</sup>Be , two <sup>14</sup>Be wavefunctions are favoured (both of which require an s-wave state near threshold in <sup>13</sup>Be as suggested by recent experiments ): the so-called D4 wavefunction โ 86% $`\nu (2s_{1/2})^2`$ and 10% $`\nu (1d_{5/2})^2`$; and C7 โ 29% $`\nu (2s_{1/2})^2`$ and 67% $`\nu (1d_{5/2})^2`$. The EMD decay energy spectra calculated for these wavefunctions for breakup at 35 MeV/nucleon on Pb are compared in figure 3 with that of the empirical Breit-Wigner deduced from the present measurements. The corresponding integrated two-neutron removal cross sections are 1.05 b (D4) and 0.395 b (C7) , compared to the measured value of 1.45$`\pm `$0.40 b. Although the strength is predicted to be concentrated at a somewhat lower energy than that observed, a large $`\nu (2s_{1/2})^2`$ admixture to the valence neutrons wavefunction is favoured. Such a result is supported by the total reaction cross section measurement of Suzuki et al. and is also in line with Lagrange mesh calculations of the <sup>14</sup>Be ground state (76% $`\nu (2s_{1/2})^2`$, 18% $`\nu (1d_{5/2})^2`$) . It should be noted that the treatment of the core as inert precludes, ab initio, the existence of any simple negative parity resonances in <sup>14</sup>Be.
Descouvemont has explored <sup>13,14</sup>Be within a microscopic cluster (<sup>12</sup>Be+n+n) model in which the core is active . In the case of <sup>13</sup>Be an s-wave state is predicted very close to threshold, whilst the energy of the d-wave resonance is well reproduced. Significantly, a strong E1 transition \[B(E1)$``$1.2e<sup>2</sup> fm<sup>2</sup>\] centred at $`E_{decay}=1.5`$ MeV is predicted in <sup>14</sup>Be, very close to the structure observed experimentally. Analysis of the corresponding energy surface suggests, however, that this transition is not associated with a true resonance . Calculations of the form of the associated continuum energy spectrum would be of considerable interest. Further support for the predictions of this model exists in the observation in a heavy-ion double charge-exchange reaction of a probable $`2^+`$ state in <sup>14</sup>Be at $`E_{decay}=0.25\pm 0.06`$ MeV , compared to a calculated value of 0.5 MeV.
It is interesting to note that the width of the structure seen in the present experiment would correspond, in the case of a true E1 resonance, to a mean lifetime ($`1/\mathrm{\Gamma }`$) of some 250$`\pm `$120 fm/c. This may be compared to a simple $`\mathrm{}\omega `$ ($`E_x=S_{2n}+E_0=3.14\pm 0.15`$ MeV) collective mode oscillation period of $``$400 fm/c, suggesting again the nonresonant nature of the observed transition.
In conclusion, the first kinematically complete breakup reaction study of <sup>14</sup>Be has been reported. Two-neutron removal cross sections, neutron angular distributions and invariant mass spectra characteristic of a halo were measured. The EMD observables indicate that the configuration of the halo neutrons contains a large $`\nu (2s_{1/2})^2`$ component. The relatively narrow structure observed near threshold in the EMD invariant mass spectrum is consistent with a soft-dipole excitation. Exploration of the continuum excitations beyond those probed here ($`E_{decay}>`$ 5 MeV) would thus be of particular interest. Additionally, spectroscopic studies of <sup>13</sup>Be and a determination of the $`\beta _2`$ of <sup>12</sup>Be, which are essential to developing more refined models describing <sup>14</sup>Be, are needed. Finally, in light of the present results, it would be highly desirable to explore the <sup>14</sup>Be continuum via other means, including inelastic scattering and surface dominated probes such as transfer or charge exchange.
The authors are grateful to the support provided by the technical and operations staff of LPC and GANIL. Discussions with Ian Thompson and Pierre Descouvemont are also acknowledged. This work was funded by the IN2P3-CNRS (France) and EPSRC (UK). Additional support was provided by the ALLIANCE programme of the British Council and the Ministรจre des Affaires Etrangรจres, and the Human Capital and Mobility Programme of the European Community (Contract no. CHGE-CT94-0056).
FIGURE CAPTIONS
Figure 1: (a) Single-neutron angular distributions for dissociation on C (open) and Pb (solid points). The results for C have been scaled by a factor of 1.8 so as to represent the nuclear contribution to dissociation on Pb (see text). (b) Deduced EMD single-neutron angular distribution for reactions on Pb.
Figure 2: Reconstructed <sup>14</sup>Be decay energy spectra for dissociation on (a) C, (b) Pb and (c) that deduced for EMD on Pb. The histogrammes correspond to the results of simulations (see text).
Figure 3: Comparison of the EMD decay energy spectra for the 3-body wavefunctions D4 and C7 with that deduced from the present experiment โ $`E_0=1.8`$ MeV and $`\mathrm{\Gamma }_0`$=0.8 MeV (shaded region). The later has been normalised to an integrated cross section of 1.45$`\pm `$0.40b.
|
warning/0006/astro-ph0006343.html
|
ar5iv
|
text
|
# Resolving the Structure of Cold Dark Matter Halos
## 1 Introduction
During the last decade there has been an increasingly growing interest in testing the predictions of variants of cold dark matter (CDM) models at subgalactic ($`100\mathrm{kpc}`$) scales. This interest was initiated by indications that observed rotation curves in the central regions of dark matter dominated dwarf galaxies are at odds with predictions of hierarchical models. Specifically, it was argued (Flores & Primack 1994; Moore 1994) that circular velocities, $`v_c(r)[GM(<r)/r]^{1/2}`$, at small galactocentric radii predicted by the models are too high and increase too rapidly with increasing radius compared to the observed rotation curves. The steeper than expected rise of $`v_c(r)`$ implies that the shape of the predicted halo density distribution is incorrect and/or that the DM halos formed in CDM models are too concentrated (i.e., have too much of their mass concentrated in the inner regions).
In addition to the density profiles, there is an alarming mismatch in the predicted abundance of small-mass ($`10^810^9h^1\text{M}\text{}`$) galactic satellites and the observed number of satellites in the Local Group (Kauffmann, White & Guiderdoni 1993; Klypin et al. 1999; Moore et al. 1999). Although this discrepancy may well be due to feedback processes (such as photoionization) which prevent gas collapse and star formation in the majority of the small-mass satellites (e.g., Bullock, Kravtsov & Weinberg 2000), the mass scale at which the problem sets in is similar to the scale in the spectrum of primordial fluctuations that may be responsible for the problems with density profiles. In the age of precision cosmology that forthcoming MAP and Planck cosmic microwave background anisotropy satellite missions are expected to bring, tests of the cosmological models at small scales may prove to be the final frontier and the ultimate challenge to our understanding of cosmology and structure formation in the Universe. However, this obviously requires detailed predictions and checks from the theoretical side, as well as higher resolution/quality observations and a good understanding of their implications and associated caveats. In this paper we focus on the theoretical predictions of the density distribution of DM halos.
A systematic study of halo density profiles for a wide range of halo masses and cosmologies was done by Navarro, Frenk & White (1996, 1997; hereafter NFW), who argued that the analytical profile of the form $`\rho (r)=\rho _s(r/r_s)^1(1+r/r_s)^2`$ provided a good description of halo profiles in their simulations for all halo masses and in all cosmologies. Here, $`r_s`$ is the scale radius which, for this profile, corresponds to the scale at which $`d\mathrm{log}\rho (r)/d\mathrm{log}r|_{r=r_s}=2`$. The parameters of the profile are determined by the haloโs virial mass $`M_{\mathrm{vir}}`$ and concentration defined as $`cr_{\mathrm{vir}}/r_s`$. NFW argued that there is a tight correlation between $`c`$ and $`M_{\mathrm{vir}}`$, which implies that the density distributions of halos of different masses can in fact be described by a one-parameter family of analytical profiles. Further studies by Kravtsov, Klypin & Khokhlov (1997), Kravtsov et al. (1998, hereafter KKBP98), Jing (2000), Bullock et al. (2000), although confirming the $`c(M_{\mathrm{vir}})`$ correlation, indicated that there is significant scatter in both the density profiles and concentrations for DM halos of a given mass.
Following the initial studies by Flores & Primack (1994) and Moore (1994), KKBP98 presented a systematic comparison of the results of numerical simulations with rotation curves of a sample of seventeen dark matter dominated dwarf and low surface brightness (LSB) galaxies.
We pointed out that the measured rotation curves of these galaxies all had the same shape with nearly linear central behavior, and furthermore, based on comparison with the density profiles of simulated halos there did not seem to be a significant discrepancy in the shape of the density profiles at the scales probed by the numerical simulations ($`0.020.03r_{\mathrm{vir}}`$, where $`r_{\mathrm{vir}}`$ is haloโs virial radius). In other words, the central density distribution in both galaxies and CDM halos was found to be shallower than $`r^1`$. These conclusions were subject to several caveats and required further testing. First, observed galactic rotation curves had to be re-examined more carefully and with higher resolution. The fact that all of the observed rotation curves used in earlier analyses were obtained using relatively low resolution HI observations required checks of the possible beam smearing effects. Also, the possibility of non-circular random motions in the central regions which could modify the rotation velocity of the gas (e.g., Binney & Tremain 1987, p. 198) had to be considered. Second, the theoretical predictions had to be tested for convergence and extended to scales $`0.01r_{\mathrm{vir}}`$.
Moore et al. (1998; see also a more recent convergence study by Ghigna et al. 1999) presented a convergence study arguing that mass resolution has a significant impact on the central density distribution of halos. They suggested that at least several million particles per halo are required to reliably model the density profiles at scales $`0.01r_{\mathrm{vir}}`$. Based on these results, Moore et al. (1998) advocated a density profile of the form $`\rho (r)(r/r_s)^{1.5}[1+(r/r_s)^{1.5}]^1`$, that behaves similarly ($`\rho r^3`$) to the NFW profile at large radii, but is steeper at small $`r`$: $`\rho r^{1.5}`$. Most recently, Jing & Suto (2000) presented a systematic study of density profiles for halo masses in the range $`2\times 10^{12}h^1\text{M}\text{}5\times 10^{14}h^1\text{M}\text{}`$. The study was uniform in mass and force resolution featuring $`510\times 10^5`$ particles per halo and force resolution of $`0.004r_{\mathrm{vir}}`$. They found that galaxy-mass halos in their simulations are well fitted by the profile<sup>2</sup><sup>2</sup>2Note that this profile is somewhat different than the profile advocated by Moore et al., but behaves similarly to the latter at small radii. Figure 9 shows that all three profiles โ NFW, Moore, and Jing & Suto โ provide good fits to dark matter halos simulated at high resolution. $`\rho (r)(r/r_s)^{1.5}[1+r/r_s]^{1.5}`$, but that cluster-mass halos are well described by the NFW profile, with logarithmic slope of the density profiles at $`r=0.01r_{\mathrm{vir}}`$ changing from $`1.5`$ for $`M_{\mathrm{vir}}10^{12}h^1\text{M}\text{}`$ to $`1.1`$ for $`M_{\mathrm{vir}}5\times 10^{14}h^1\text{M}\text{}`$. Jing & Suto interpreted these results as an evidence that profiles of DM halos are not universal (but see ยง3.1 for a possible alternative interpretation).
At small scales, the results of Kravtsov et al. (1998) are at odds with the results of above studies. Although fairly extensive convergence tests were done in that study, they focused on the effects of spatial resolution and the mass resolution was kept constant in almost all the tests. In this case we found that halo density profiles converged at scales larger than two formal resolutions of the ART code. As we will show in this paper, this is not true when the mass resolution is varied. In particular, the convergence study presented in this paper in which we varied both mass and force resolution shows that for the ART simulations convergence is reached at scales larger than four formal resolutions or the scales containing 200 particles, whichever is larger. The shallow behavior of the density profiles in Kravtsov et al. was found at scales $`25`$ formal resolutions (at larger scales profiles were consistent with the NFW functional form) and is therefore a numerical artifact. In simulations presented in this paper we find profiles that are consistent with cuspy NFW and Moore et al. distributions at well resolved scales.
New observational and theoretical developments show that comparison between model predictions and observational data is not straightforward. Decisive comparisons require reaching convergence of theoretical predictions and understanding the kinematics of the gas in the central regions of observed galaxies. As we noted above, in this paper we present convergence tests designed to test effects of mass resolution on the density profiles of halos formed in the currently popular CDM model with cosmological constant ($`\mathrm{\Lambda }`$CDM) and simulated using the multiple mass resolution version of the Adaptive Refinement Tree code (ART). This study is crucial in resolving the discrepancy of our previous study on the density profiles with other numerical studies. We also discuss several caveats with respect to drawing conclusions about the density profiles from the fits of analytical functions to numerical results and their comparisons to observational data. In the following section we describe the code and numerical simulations used in our analysis. In ยง3 we compare the analytical fits advocated by NFW and Moore et al., fits of these profiles to the density profiles of simulated halos, and convergence analysis of our numerical results.
## 2 Numerical simulations
### 2.1 Code description
The Adaptive Refinement Tree code (ART; Kravtsov, Klypin & Khokhlov 1997) was used to run the simulations. The ART code starts with a uniform grid, which covers the whole computational box. This grid defines the lowest (zeroth) level of resolution of the simulation. The standard Particles-Mesh algorithms are used to compute density and gravitational potential on the zeroth-level mesh. The ART code reaches high force resolution by refining all high density regions using an automated refinement algorithm. The refinements are recursive: the refined regions can also be refined, each subsequent refinement having half of the previous levelโs cell size. This creates a hierarchy of refinement meshes of different resolution, size, and geometry covering regions of interest. Because each individual cubic cell can be refined, the shape of the refinement mesh can be arbitrary and match effectively the geometry of the region of interest.
The criterion for refinement is the local density of particles: if the number of particles in a mesh cell (as estimated by the Cloud-In-Cell method) exceeds the level $`n_{\mathrm{thresh}}`$, the cell is split (โrefinedโ) into 8 cells of the next refinement level. The refinement threshold may depend on the refinement level. The code uses the expansion parameter $`a`$ as the time variable. During the integration, spatial refinement is accompanied by temporal refinement. Namely, each level of refinement, $`l`$, is integrated with its own time step $`\mathrm{\Delta }a_l=\mathrm{\Delta }a_0/2^l`$, where $`\mathrm{\Delta }a_0`$ is the global time step of the zeroth refinement level. This variable time stepping is very important for accuracy of the results. As the force resolution increases, more steps are needed to integrate the trajectories accurately. In the remainder of the paper by the term formal resolution, $`h_{\mathrm{formal}}`$, we will mean the size of a cell on the highest level of refinement reached in simulation. This is similar to the usual practice of defining formal resolution in the uniform grid codes. The actual force resolution of the code is somewhat larger than the formal resolution. In the ART code, the interparticle force is weaker (โsofterโ) than the Newtonian force at scales $`<2h_{\mathrm{formal}}`$. The average interparticle force is Newtonian at scales $`\begin{array}{c}>\hfill \\ \hfill \end{array}2h_{\mathrm{formal}}`$, although there is a substantial scatter in the force at $`23h_{\mathrm{formal}}`$ due primarily to errors in numerical differentiation of potential. For comparison, average Plummer softened force with softening $`ฯต_{\mathrm{Plummer}}=h_{\mathrm{formal}}`$, reaches the Newtonian value at scales $`56h_{\mathrm{formal}}`$ (see Kravtsov et al. 1997 for details). Therefore, one formal resolution in the ART code is equivalent to $`0.30.4`$ Plummer softening of the same value.
The cosmological simulations performed using the ART code were compared with the simulations started from identical initial conditions and performed using the well-known PM and AP<sup>3</sup>M codes. The comparisons showed that results of ART simulations (for a wide battery of the commonly used statistics and halo parameters) are similar to those of the AP<sup>3</sup>M simulations at all resolved scales. These comparisons and other tests of the ART code can be found in Kravtsov (1999) and Knebe et al. (2000).
### 2.2 Initial conditions
The current version of the ART code has the ability to handle particles of different masses. In the present analysis this ability was used to increase the mass (and correspondingly the force) resolution inside a few pre-selected halos. The multiple mass resolution is implemented in the following way. We set up a realization of the initial spectrum of perturbations in such a way that a very large number of small-mass particles can be generated in the simulation box. For example, for the first (second) set of simulations (see below) we generate $`512^3`$ ($`1024^3`$) independent spectrum harmonics. Potentially, initial conditions with $`512^3/1024^3`$ particles could be generated. Coordinates and velocities of the particles are calculated using all waves ranging from the fundamental mode $`k=2\pi /L`$ to the Nyquist frequency $`k=2\pi /L\times N^{1/3}/2`$, where $`L`$ is the box size and $`N`$ is the number of particles in the simulation.
The code actually generates positions and velocities for all $`512^3/1024^3`$ particles, but some of the particles are then merged into particles of larger mass. The larger mass (merged) particle is assigned velocity and displacement equal to the average velocity and displacement of the merged particles. The whole lagrangian space of particles is divided into large cubic blocks of particles with each block having $`16^3`$ particles. Depending on what local mass resolution is required, each particular block can be subdivided into smaller sub-blocks and generate from 1 to $`16^3`$ particles (the highest resolution). Using this procedure<sup>3</sup><sup>3</sup>3 The code is actually written to handle an arbitrary dynamic range. The current limit is determined by computational limitations., we can generate particles with 5 different masses covering dynamic mass range of 4096.
We start simulations by making a low resolution run with uniform mass resolution in which all particles have the largest possible mass. Next we run simulations with $`32^3`$ and $`64^3`$ particles. Using these runs, we identify halos in the simulation and select halos to be re-simulated with higher mass and force resolution. For each selected halo we determine its virial radius. We then identify all particles inside the two virial radii and find lagrangian coordinates of each particle. The coordinates are used to mark blocks of particles to generate the initial conditions of the highest mass resolution. Once all particles are processed and all blocks are marked, we mark all blocks adjacent to those already marked to produce initial conditions of the eight times lower mass resolution. This procedure is repeated for lower and lower mass resolution levels. In the end, each unmarked block will produce one most massive particle and a marked block will generate a number of particles which depends on the step in which the block was marked. Figure 1 shows the outcome of the process of mass refinement in a 2-dimensional case.
Figure 2 shows an example of mass refinement for one of the halos in our simulations. A large fraction of high resolution particles ends up in the central halo, which does not have any larger mass particles (see insert in the bottom panel). At $`z=10`$, the region occupied by the high resolution particles is non-spherical: it is substantially elongated in the direction perpendicular to the large filament clearly seen at $`z=0`$.
After the initial conditions are set, we run the simulation again allowing the code to perform mesh refinement based only on the number of particles with the smallest mass.
### 2.3 Numerical simulations
We simulated a flat low-density cosmological model ($`\mathrm{\Lambda }`$CDM) with $`\mathrm{\Omega }_0=1\mathrm{\Omega }_\mathrm{\Lambda }=0.3`$, the Hubble parameter (in units of $`100\mathrm{kms}^1\mathrm{Mpc}^1`$) $`h=0.7`$, and the spectrum normalization $`\sigma _8=0.9`$. We have run two sets of simulations. The first set used $`128^3`$ zeroth-level grid in a computational box of $`30h^1\text{Mpc}`$. The second set of simulations used $`256^3`$ grid in a $`25h^1\text{Mpc}`$ box and had higher mass resolution. In the simulations used in this paper, the threshold for cell refinement (see above) was low on the zeroth level: $`n_{\mathrm{thresh}}(0)=2`$. Thus, every zeroth-level cell containing two or more particles was refined. This was done to preserve all small-scale perturbations present in the initial spectrum of perturbations. The threshold was higher on deeper levels of refinement. For the first set of simulations it was $`n_{\mathrm{thresh}}=2`$ at the first refinement level and $`n_{\mathrm{thresh}}=3`$ for all higher levels. For the second simulation the thresholds were $`n_{\mathrm{thresh}}=3`$ and $`n_{\mathrm{thresh}}=4`$ for the first level and higher levels, respectively.
In addition to effects introduced by limited mass and force resolution, integration errors of particle trajectories may affect the innermost regions of halos. The local dynamical time for particles moving in these regions is quite short. For example, the period of a particle on a circular orbit of radius $`1h^1\text{kpc}`$ around the center of halo A is only $`0.5\%`$ of the Hubble time. Therefore, if the time step is not sufficiently small, numerical errors in these regions will tend to grow. Even for small time steps errors exist and tend to alter the density distribution in the centers of halos over some limited range of scales.
All of our simulations were started at $`z_i=60`$ and the step in the expansion parameter was chosen to be $`\mathrm{\Delta }a_0=2\times 10^3`$ for particles located on the zeroth base grid. This gives about 500 steps for particles located in the zeroth level for an entire run to $`z=0`$. We have done a test run with twice smaller time step for a halo of mass comparable (but with smaller number of particles) to the mass of halos studied in this paper. We did not find any significant differences in the resulting halo profile. For both sets of simulations, the highest level of refinement was ten for the largest mass resolution, which corresponds to $`500\times 2^{10}500,000`$ time steps at the tenth refinement level. Some simulations were rerun with smaller number of particles. They did not reach the highest levels of refinement, and, thus, they had fewer steps. For example, halo D<sub>2</sub> has reached only 7 levels of refinement and had only $`500\times 2^764,000`$ time steps.
In the following sections we present density profiles of four halos. The halo A was the only halo selected for re-simulation in the first set of simulations. It was relatively quiescent at $`z=0`$ and had no massive neighbors. The halo was located in a long filament bordering a large void and was about 10 Mpc away from the nearest cluster-size halo. After the high-resolution simulation was completed we found that the nearest galaxy-size halo was about 5 Mpc away. The halo had a fairly typical merging history with $`M(t)`$ track slightly lower than the average mass growth predicted using the extended Press-Schechter model. The last major merger event occurred at $`z2.5`$; at lower redshifts the mass growth (the mass in this time interval has grown by a factor of three) was due to slow and steady mass accretion.
The halos B, C, and D were identified in the second set of simulations and were selected among halos residing in a well defined filament. Two of the halos (B and C) are neighbors located about 0.5 Mpc from each other. The third halo was 2 Mpc away from this pair. Thus, the halos were not selected to be too isolated as was the case in the first set of runs. Moreover, the simulation was also analyzed at both $`z=0`$ and at $`z=1`$ (when halos are more likely to be less relaxed). Therefore, the halo A can be considered as an example of a rather isolated well-relaxed halo. In many respects, this halo is similar to halos simulated by other research groups that used multiple mass resolution techniques. The halos B, C, and D from the second set of simulations can be viewed as representative of more typical halo population located in more crowded environments.
Parameters of the simulated dark matter halos are listed in Table 1. Columns in the table present (1) haloโs โnameโ (halos A<sub>1</sub>, A<sub>2</sub>, A<sub>3</sub> are the halo A re-simulated three times with different mass and force resolutions); (2) redshift at which the halo was analyzed; (3) the number of particles within the virial radius; (4) the smallest particle mass in the simulation; (5) formal comoving force resolution (cells size at the highest refinement level) achieved in the simulation.
## 3 Results
### 3.1 Comparison of the NFW and the Moore et al. profiles
Before we fit analytical profiles to profiles of simulated dark matter halos or compare them to the observed rotation curves, it is instructive to compare different analytical approximations. Although the NFW and Moore et al. profiles predict different behavior of $`\rho (r)`$ in the central regions of a halo, the scale at which this difference becomes significant depends on the specific values of the haloโs characteristic density and radius. Table 2 presents the parameters and statistics associated with the two analytical profiles. For the NFW profile more information can be found in Klypin et al. (1998), ลokas & Mamon (2000), and Widrow (2000).
Each profile is defined by two independent parameters. We choose these to be the characteristic density $`\rho _s`$ and radius $`r_s`$. In this case all expressions describing the properties of the profiles have a simple form and do not depend on the concentration. Both the concentration and the virial mass appear only in the normalization of the expressions. The choice of the virial radius (e.g., ลokas & Mamon 2000) as a scale unit results in more complicated expressions with explicit dependence on the concentration. In this case, one has to be careful about the definition of the virial radius, as there are several definitions in the literature. For example, it is often defined as the radius, $`r_{200}`$, within which the average density is 200 times the critical density. In this paper the virial radius is defined as the radius within which the average density is equal to the density predicted by the top-hat model: it is $`\delta _{\mathrm{TH}}`$ times the average matter density in the Universe. In the case of $`\mathrm{\Omega }_0=0.3`$ models the virial radius defined in this way is about 30% larger than $`r_{200}`$ (e.g., Eke et al. 1998).
There is no unique way of defining a consistent concentration for the different analytical profiles. Again, it is natural to use the characteristic radius $`r_s`$ to define the concentration: $`cr_{\mathrm{vir}}/r_s`$. This simplifies the expressions. At the same time, if we fit the dark matter halo with the two profiles, we will get different concentrations because the values of the corresponding $`r_s`$ will be different. Alternatively, if we choose to match the outer regions of the profiles (say, $`r>r_s`$) as closely as possible, we may choose to change the ratio of the characteristic radii $`r_{s,\mathrm{NFW}}/r_{s,\mathrm{Moore}}`$ in such a way that both profiles reach the maximum circular velocity $`v_{\mathrm{circ}}`$ at the same physical radius $`r_{\mathrm{max}}`$. In this case, the formal concentration of the Moore et al. profile is 1.72 times smaller than that of the NFW profile. Indeed, with this normalization profiles look very similar in the outer parts as one finds in Figure 3. Table 2 also gives two other โconcentrationsโ. The concentration $`C_{1/5}`$ is defined as the ratio of virial radius to the radius, which encompasses 1/5 of the virial mass (Avila-Reese et al. 1999). For halos with $`C_{\mathrm{NFW}}5.5`$ this 1/5 mass concentration is equal to $`C_{\mathrm{NFW}}`$. One can also define the concentration as the ratio of the virial radius to the radius at which the logarithmic slope of the density profile is equal to $`2`$. This scale corresponds to $`r_s`$ for the NFW profile and $`0.35r_s`$ for the Moore et al. profile.
Figure 3 presents the comparison between the analytic profiles normalized to have the same virial mass and the same radius $`r_{\mathrm{max}}`$. We show results for halos of low and high values of concentration representative of cluster- and low-mass galaxy halos, respectively. The bottom panels show the profiles, while the top panels show the corresponding logarithmic slope as a function of radius. The figure shows that the two profiles are very similar throughout the main body of the halos. Only in the very central region do the differences become significant. The difference is more apparent in the logarithmic slope than in the actual density profiles. Moreover, for galaxy-mass halos the difference sets in at a rather small radius $`0.01r_{\mathrm{vir}}`$, which would correspond to scales $`<1\mathrm{kpc}`$ for the typical dark matter dominated dwarf and LSB galaxies. At the observationally interesting scales the differences between NFW and Moore et al. profiles are fairly small and the NFW profile provides an accurate description of the halo density distribution.
Note also that for galaxy-size (e.g., high-concentration) halos the logarithmic slope of the NFW profile has not yet reached its asymptotic inner value of $`1`$ even at scales as small as $`0.01r_{\mathrm{vir}}`$. At this distance the logarithmic slope of the NFW profile is $`1.41.5`$ for halos with mass $`10^{12}h^1\text{M}\text{}`$. For cluster-size halos this slope is $`1.2`$. This dependence of the slope at a given fraction of the virial radius on the virial mass of the halo is very similar to the results plotted in Figure 3 of Jing & Suto (2000). These authors interpreted it as evidence that halo profiles are not universal. It is obvious, however, that their results are consistent with NFW profiles and the dependence of the slope on mass can be simply a manifestation of the well-studied $`c_{\mathrm{vir}}(M)`$ relation.
The NFW and Moore et al. profiles can be compared in a different way. We can approximate the Moore et al. halo of a given concentration with the NFW profile. Fractional deviations of the fits depend on the halo concentration and on the range of radii used for the fits. A low-concentration halo has larger deviations, but even for $`C=7`$ case, the deviations are less than 15% if we fit the halo at scales $`0.01<r/r_{\mathrm{vir}}<1`$. For a high-concentration halo with $`C=17`$, the deviations are much smaller: less than 8% for the same range of scales.
To summarize, we find that the differences between the NFW and the Moore et al. profiles are very small ($`\mathrm{\Delta }\rho /\rho <10\%`$) for radii above 1% of the virial radius for typical galaxy-size halos with $`C_{\mathrm{NFW}}\begin{array}{c}>\hfill \\ \hfill \end{array}12`$. The differences are larger for halos with smaller concentrations. In the case of the NFW profile, the asymptotic value of the central slope $`\gamma =1`$ is not achieved even at radii as small as 1%-2% of the virial radius.
### 3.2 Convergence study
The effects of numerical resolution can be studied by resimulating the same objects with higher force and mass resolution and with a larger number of time steps. In this study we performed simulations of the same halos with increasingly higher mass resolution. In the ART code simulations the subsequent mesh refinements are done when particle density in a mesh cell exceeds a specified threshold. The mass resolution is thus tightly linked with the highest achievable spatial resolution.
Table 3 gives parameters of the halos and parameters of their fits. The first and the second columns give the halo name (Table 1) and the redshift at which the halo was studied. Columns (3-5) present virial mass, radius, and the maximum circular velocity of the halo. Columns (6-8) present parameters of the fits: the halo concentration as estimated using the NFW profile and the maximum relative errors of the NFW and Moore et al fits. The bottom panel in figure 4 shows density profiles for the simulations of halo A (see Table 1). Here, as in the Fig.1a in Moore et al. (1998), all profiles are plotted down to the formal force resolution of the corresponding run. Although it may appear that density profiles have not converged in the central region and that the low resolution simulations produce erroneous results, this is simply an artifact of plotting the profiles below the actual numerical resolution (or convergence scale). The top panel in figure 4 and figure 5 show profiles of halos A, B, C, and D plotted down to four formal resolutions of the simulations (4 mesh cells at the highest refinement level). The figures show that in this case density profiles in lower resolution simulations are in very good agreement with profiles in high-resolution runs at all radii. For example, there are no systematic differences in the logarithmic slope of the profile at a given distance and we find no significant change in the concentration parameters or the maximum circular velocities of halos (see Table 3).
Closer examination of the bottom panel in Figure 4, shows that profiles have not converged at two formal resolutions. This is at odds with our convergence study in Kravtsov et al. (1998). We attribute this difference to the fact that mass resolution in the latter study was kept fixed when force resolution was varied. Our highest resolution run A<sub>1</sub>, if considered including scales larger than two formal resolutions, is consistent with conclusion about the shallow central slope made in Kravtsov et al. (1998). Indeed, if the profiles are considered down to the scale of two formal resolutions (a scale smaller than the smallest converged scale), the density profile slope in the very central part of the profile $`r0.01r_{\mathrm{vir}}`$ is close to $`\gamma =0.5`$. In light of the results shown in Figure 4, it is clear that this is an artifact of underestimating true convergence scale and conclusions about shallow central density distributions made in Kravtsov et al. (1998) are therefore incorrect.
It is not clear which numerical effect determine the convergence scale. It is likely that this scale is determined by a complex interplay of all numerical effects. For the ART simulations we found empirically that the scale above which the density does not deviate (deviations were less than 10%) from results of higher resolution simulation is $`4`$ formal force resolutions or containing more than 200 particles, whichever is larger. The limit very likely depends on particular code used and is not universal. Figure 4, top panel, shows that for the halo A, convergence for vastly different mass and force resolution is reached for scales $`4`$ formal force resolutions (all profiles in this figure are plotted down to the radius of 4 formal force resolutions). For all resolutions, there are more than 200 particles within the radius of four resolutions from the halo center. For the highest resolution simulation (halo A<sub>1</sub>) convergence is reached at scales $`0.005r_{\mathrm{vir}}`$, assuming convergence at 4 times the formal resolution as found for halos A<sub>2</sub> and A<sub>3</sub>. For halos B, C, D (figure 5) this criterion also worked, but was mostly defined by the number of particles (more than 200-300 particles for convergence).
### 3.3 Halo profiles
In order to judge which analytical profile provides a better description of the simulated profiles we fitted the NFW and Moore et al. analytic profiles. Figure 6 presents results of the fits for halo A and shows that both profiles fit the simulated profile equally well: fractional deviations of the fitted profiles from the numerical one are smaller than 20% over almost three decades in radius. It is thus clear that the fact that the numerical profile has slope steeper than $`1`$ at the scale of $`0.01r_{\mathrm{vir}}`$ does not mean that a good fit of the NFW profile (or even analytic profiles with shallower asymptotic slopes) cannot be obtained. Figure 7 shows the fitting of halos in the second set of simulations. Each halo in the plot has more than a million particles โ ten times more than halo A. One would naively expect that this increase in the resolution should clearly show which profile makes a better fit. Indeed, more particles and better resolution gave smaller deviations, but the fits became better for both approximations. For example, at 1% of the virial radius of the halo D the deviations were 3.6% for the NFW profile and 6.2% for the Moore et al. profile โ down from 20% for the halo A at the same distance. The Moore et al. approximation gave a better fit for halos B and C, but not for halo D. The NFW approximation was less accurate on intermediate scales around $`(0.030.1)R_{\mathrm{vir}}`$, but the errors were quite small. Thus, both approximations gave comparable results.
There is definitely a certain degree of degeneracy in fitting various analytic profiles to numerical results. Figure 8 illustrates this further by showing results of fitting profiles (solid lines) of the form $`\rho (r)(r/r_0)^\gamma [1+(r/r_0)^\alpha ]^{(\beta \alpha )/\gamma }`$ to the same simulated halo profile (halo A<sub>1</sub>) shown as solid circles. The legend in each panel indicates the corresponding values of $`\alpha `$, $`\beta `$, and $`\gamma `$ of the fit; the digit in parenthesis indicates whether the parameter was kept fixed ($`0`$) or not ($`1`$) during the fit. The two right panels show fits of the NFW and Moore et al. profile; the bottom left panel shows fit of the profiles used by Jing & Suto (2000). The top left panel shows a fit in which the inner slope was fixed but $`\alpha `$ and $`\beta `$ were fit. The figure shows that all four analytic profiles can provide a good fit to the numerical profile in the whole range of resolved scales: $`0.0051r_{\mathrm{vir}}`$.
As we mentioned in ยง 2.3, the halo A analyzed in the previous section is somewhat special because it was selected as an isolated relaxed halo. Halos, which are not very isolated and relaxed are also interesting. After all, they represent the majority of all halos. When we compare observed rotation curves of galaxies with predicted circular velocity curves of the halos, we do not know if the galaxy host halo is well relaxed or not. In order to reach unbiased conclusions, we will present analysis of halos from the second set of simulations at redshifts $`z=0`$ and $`z=1`$ (halos B<sub>1,3</sub>, C<sub>1,3</sub>, and D<sub>1,3</sub>), which were not selected to be relaxed or isolated. Note that these halos did not have major mergers immediately prior to the epoch of analysis, which could produce large distortions of their profiles. Based on the results of the convergence study presented in the previous section, we will consider profiles of these halos only at scales above four formal resolutions and not less than 200 particles. There is an advantage in analyzing halos at a relatively high redshift. Halos of a given mass will have lower concentration (see Bullock et al. 2000). Lower concentration implies a large scale at which the asymptotic inner slope is reached.
We found that substantial substructure is present inside the virial radius in all three halos at $`z=1`$. Figure 7 shows profiles of these halos at $`z=0`$ (top) and $`z=1`$ (bottom). The $`z=0`$ profiles are smoother than profiles at $`z=1`$. Note that bumps and depressions visible in the profiles have amplitude that is significantly larger than the shot noise. Halo C<sub>3</sub> appeared to be the most relaxed of the three halos. This halo had its last major merger somewhat earlier than the other two. Halo D<sub>3</sub> had a major merger event at $`z2`$. A remnant of the merger is still visible as a bump at $`r100h^1\text{kpc}`$. The non-uniformities of profiles caused by substructure may substantially bias analytic fits if one uses the entire range of scales below the virial radius. Therefore, we used only the central, presumably more relaxed, regions in the analytic fits: $`r<50h^1\text{kpc}`$ for halo D and $`r<100h^1\text{kpc}`$ for halos B and C (fits using only central $`50h^1\text{kpc}`$ did not change results).
The best fit parameters were obtained by minimizing the maximum fractional deviation of the fit: $`\mathrm{max}[\mathrm{abs}(\mathrm{log}\rho _{\mathrm{fit}}\mathrm{log}\rho _{\mathrm{halo}})]`$. Minimizing the sum of squares of deviations ($`\chi ^2`$), as is often done, can result in larger errors at small radii with the false impression that the fit fails because it has a wrong central slope. The fit that minimizes maximum deviations improves the NFW fit for points in the range of radii $`(520)h^1\text{kpc}`$, where the NFW fit would appear to be below the data points if the fit was done by the $`\chi ^2`$ minimization. For example, if we fit halo B by minimizing $`\chi ^2`$, the concentration slightly decreases from 12.3 (see Table 1) to 11.8, the maximum error slightly increases to 27%, but the fit goes below the data points for most of the points at small radii.
We have also fitted density distribution of halo B assuming even more stringent limits on the effects of numerical resolution. We fitted the halo starting at the scale equal to six times the formal resolution, minimizing the maximum deviation. Inside this radius there were about 900 particles. Resulting parameters of the fit were close to those in Table 1: $`C_{\mathrm{NFW}}=11.8`$, and maximum error of the NFW fit was 17%.
We found that for halos B and C the errors in the Moore et al. fits were systematically smaller than those of the NFW fits, though the differences were not dramatic. But Moore et al. fit poorly in the case of halo D. It formally gave very small errors, but at the expense of unreasonably small concentration $`C_{\mathrm{NFW}}=2`$. When we constrained the approximation to have about twice larger concentration as compared with the best NFW fit, we were able to obtain a reasonable fit (this fit is shown in Figure 7). Nevertheless, the central density distribution is fit poorly in this case.
Therefore, our analysis does not show that one analytic profile is better then the other for description of the density distribution in simulated halos. Despite the larger number of particles per halo and the lower concentrations of $`z=1`$ halos, results are still inconclusive. The Moore at al. profile is a better fit to the profile of halo C; the NFW profile is a better fit to the central part of the halo D. Halo B represents an intermediate case where both profiles provide equally good fits (similar to the analysis of halo A). Remarkably, the same conclusions hold for the halo profiles at $`z=0`$.
Both at $`z=0`$ and $`z=1`$, there are real deviations in parameters of halos of the same mass. We find the same differences in estimates of $`C_{1/5}`$ concentrations, which do not depend on specifics of an analytic fit. The central slope at around $`1\text{kpc}`$ also changes from halo to halo. Halos B and C have the same virial radii and nearly the same circular velocities, yet their concentrations are different by 30%. Indeed, halos in the Table 3 have similar masses in the range $`(1.22)\times 10^{12}h^1\text{M}\text{}`$. If halos had a universal profile โ a shape, which depends only on halo mass, then we should expect that the circular velocity curves are very similar for our halos. Figure 9 shows circular velocities for halos B, C, and D, which have only 25% deviations in their virial mass. The halos clearly do not have a universal one-parameter shape. There are substantial variations in the curves, which occur at relatively large radii ($`0.10.3R_{\mathrm{vir}}`$). The variations are due to differences in halo concentration โ each curve is well described by a NFW or Moore et al. profile but their concentrations are somewhat different. Our three halos clearly constitute a small sample. Bullock et al. (2000) and Jing (2000) studied the spread of halo concentrations in a large sample of halos. For a given mass it was found that halos have 20-50% variations in the concentration at $`1\sigma `$ level, which is consistent with what we find for our halos.
## 4 Discussion and conclusions
We have analyzed a series of simulations with vastly different mass and force resolutions with the goal of studying density distribution in the central regions of galaxy-size dark matter halos. We used multiple mass simulations performed using the ART code; the simulations were performed with variable mass, force and temporal resolutions. In the highest resolution runs, we achieved a (formal) spatial dynamical range of $`2^{18}=262144`$; the simulation was run with 500,000 steps for particles at the highest level of refinement.
Using these simulations, we have studied convergence of halo density profiles for different mass and force resolutions. We show that the halo profiles converge at scales larger than a certain (true numerical resolution) scale defined by numerical effects. This scale is probably code dependent, but can be found for any numerical code by a convergence study. For the ART simulations presented here, the density profiles converged at the scale of four times the formal force resolution or the radius containing more than 200 particles, whichever is larger. In this sense, our results are consistent with results of the โSanta Barbaraโ cluster comparison project (Frenk et al. 1999): the density profiles of a cluster-size halo simulated with different numerical codes and resolutions agree with each other at all resolved scales.
In KKBP98 we have discussed convergence tests in which we varied force resolution while keeping mass resolution fixed. Using these tests we concluded that density profiles converge at radii twice the local formal resolution of a simulation. The convergence tests presented in this paper, however, show that mass resolution places more stringent conditions on the trustworthy range of scales. Although we can reproduce our previous results (shallower than $`1`$ density profiles at radii of two formal resolution), our new convergence tests show that these results were affected by limited mass resolution. We conclude that we overestimated our force resolution KKBP98 and that the conclusions about the shallow central slopes presented there were an artifact of that overestimate. Other results in KKBP98 that focus on general halo characteristics, such as the scatter in profile shapes, and the agreement between the $`V_{max}`$$`r_{max}`$ relations of simulated dark halos and dark matter dominated dwarf and LSB galaxies, are valid.
At scales above four times the formal resolution and containing more than 200 particles results presented in this paper agree well with previous simulations. For example, the concentration parameter for the halos are in good agreement with the concentration-mass dependence ($`c(M)`$) presented in Bullock et al. (2000) based on the previous simulations.
We can also reproduce results of convergence studies by Moore et al. (1998) and Ghigna et al. (1999). At first glance it may appear that our conclusions are in direct contradiction with these studies that concluded that at least several millions of particles are needed to resolve a halo profile properly. However, the contradiction is only in interpretation rather than in the results themselves. For example, Figure 2 in Ghigna et al. shows a cluster profile simulated at 3 different mass and force resolutions. The conclusion the authors make based on this convergence study is (at least qualitatively) similar to our conclusion: the profiles converge at all mass resolutions at scales above $`6ฯต`$, where $`ฯต`$ is the spline force softening of their code<sup>4</sup><sup>4</sup>4Note that profiles of lower resolution simulations perfectly agree with profile of the highest resolution run at scales above $`34ฯต`$. (roughly equivalent to our formal resolution). Why is this criterion is more severe than ours? The obvious reasons are differences in the force shape and other differences between numerical codes. However, it appears that the softening in these simulations was set too low (the profiles were โoverresolvedโ) and convergence scale is determined by the number of particles criterion rather than by force resolution. Indeed, we can estimate the number of particles within the softening scale in these simulations as $`N_p(<ฯต)(4\pi /3)(\rho _{crit}/m_p)(1+\delta )ฯต^3,`$ where $`ฯต`$ is the softening length, $`m_p`$ is the particle mass, $`\rho _{crit}`$ is the critical density, and $`\delta `$ is the overdensity reached at the softening scales. For the HIRES simulation in Ghigna et al., $`ฯต=0.5h^1\mathrm{kpc}`$ and $`m_p=5.37\times 10^7h^1\mathrm{M}_{}`$ and $`\delta (0.5รท1.0)\times 10^7`$, which gives $`N_p(<ฯต)1030`$. For their LOWRES simulation this number is even lower $`510`$. The scale containing $`200300`$ particles should be close to the convergence that is found in this study (e.g., $`34ฯต`$ for the LOWRES run).
In light of these considerations, the profiles in Fig 2 of Moore et al. (1998) are perfectly consistent with each other if considered at scales $`4ฯต`$. Their results are thus perfectly consistent with our results and conclusions. One needs more particles only if the softening is set too small and the inner regions are over-resolved. In this case the true resolution is set by the radius that contains at least a couple hundred particles. The main point is that the higher mass resolution is needed to probe deeper into the inner regions of halos. However, if one is interested only in the profiles at, say, $`r0.02R_{vir}`$ (sufficient to determine haloโs concentration, maximum circular velocity, etc.) than, as Figure 2 in Ghigna et al. (1999) and Figure 5 in this paper clearly show, $`(25)\times 10^5`$ within the virial radius is adequate.
In this paper, we present results for halos that contain $`1.2\times 10^5`$ to $`1.6\times 10^6`$ particles within their virial radius. We show that for the galaxy-size halos ($`M_{\mathrm{vir}}=7\times 10^{11}h^1\text{M}\text{}2\times 10^{12}h^1\text{M}\text{}`$ and $`C=917`$) both the NFW profile $`\rho r^1(1+r)^2`$ and the Moore et al. profile $`\rho r^{1.5}(1+r^{1.5})^1`$ provide fairly good fits of the simulated profiles with deviations of about 10% for radii larger than 1% of the virial radius. For dwarf and LSB galaxies commonly used for comparisons with model predictions, this corresponds to scales $`\begin{array}{c}>\hfill \\ \hfill \end{array}1`$kpc. Therefore, the debate about which analytic profile provides a better description of the CDM halo profiles may well be irrelevant for comparisons to measured galaxy rotation curves. Such comparisons are also subject to other uncertainties, one of which is limited spatial extent of the observed rotation curves. The particular shape of the inner density distribution may be important for galaxy cluster observations, however. Cluster-size halos are predicted to have smaller concentrations, which means that the scale at which differences between the NFW and Moore et al. profiles become significant is larger and is observationally relevant.
These results are consistent with results of Moore et al. (1999) who found that galaxy-size halos in CDM models are well described by the $`\rho r^{1.5}(1+r^{1.5})^1`$ profile. The authors used simulations of the standard CDM model with mass and force resolution similar to our simulations ($`2\times 10^6M_{}`$ and $`1`$ kpc, respectively). They found that the NFW profile that was fitted at radii above $`3\%`$ of the haloโs virial radius, underpredicts the density at smaller radii by up to $`2030\%`$. This is consistent with out results for halos B and C, although not for halo D, which probably indicates a certain degree of variance among profiles of halos of the same mass (Jing 2000; Bullock et al. 2000; Avila-Reese et al. 1999).
Jing & Suto (2000) simulated formation of galaxy-, group-, and cluster-size halos with resolution similar to our highest-resolution simulations. They also found that the NFW profile fit to the outer regions (radii above few percent of the virial radius) underestimates the density in the innermost regions of the halo. The degree of discrepancy, however, appears to be different for different halos (see their Fig. 2), which is consistent with our conclusions. The authors conclude that the shapes of the density profiles vary from galaxy- to group-, to cluster-size halos. However, as we argued in ยง 3.1, their results could be interpreted as the manifestation of concentration-mass relation instead.
Most recently, Fukushige & Makino (2001) simulated 12 halos with masses ranging from $`4.3\times 10^{11}M_{}`$ to $`7.9\times 10^{14}M_{}`$, again with resolution similar to the resolution of simulations presented here. The dependence of the inner logarithmic slope on the halo mass observed by Jing & Suto (2000) was not found in these simulations. Instead, the steepest slope of the density profiles was found to be close to $`1.5`$ for all masses. This is surprising because such dependence is expected from the concentration-mass correlation (e.g., NFW; Bullock et al. 2000). It is not clear what could explain this discrepancy. Note also that setup of these simulations was somewhat different from the setup that is usually used: the simulations followed halo collapse from a spherically symmetric configuration centered on a gaussian density peak with no external tidal field included.
We show that density profiles of halos that are not fully relaxed may contain real non-uniformities due to substructure and differences in the density distributions. These non-uniformities affect the fit quality for a particular analytic profile and result in somewhat different values of fitted parameters (e.g., concentration). This was clearly seen at redshift $`z=1`$ for halos B,C,D (Figure 7), which by that time have not yet relaxed. At redshift $`z=0`$ the halos were much more quiet and the deviations from the fits were much smaller. It is interesting to note that the non-equilibrium effects do not qualitatively change the shape of the central parts of density profiles. For example, we found that at both redshifts the profile of halo C is best fit by the Moore et al. profile. The density profile of halo D, however, is best fit by the NFW profile again at both redshifts. Note that these three halos were simulated with the same mass and force resolution (indeed, in the same simulation). It seems that the main difference between these halos is their merger histories. We conclude, therefore, that differences in merger history and/or different degree of substructure in halos of the same mass may explain the scatter in profile shapes and concentration parameters found in previous studies (KKBP98; Jing 2000; Bullock et al. 2000).
In view of the current less confusing situation regarding the theoretical predictions for CDM profiles, it is interesting to discuss how the theory compares to observations. Rotation curves of a number of dwarf and LSB galaxies have recently been re-examined using H$`\alpha `$ observations and/or including corrections for beam-smearing in HI observations (e.g., Swaters, Madore & Trewhella 2000; van den Bosch et al. 2000). The results show that for the majority of galaxies, the H$`\alpha `$ rotation curves are significantly different in their central regions than the rotation curves derived from HI observations. This indicates that the HI rotation curves are affected by beam smearing (Swaters et al. 2000). This also implies that beam smearing may be at least partly responsible for the universal shape of the LSB rotation curves discussed in KKBP98.
It is possible that part of the discrepancy between the rotation curves of the different tracers may be due to real differences in the kinematics of the two gas components (ionized and neutral hydrogen). Preliminary comparisons between the new H$`\alpha `$ rotation curves and model predictions show that NFW density profiles are consistent with the observed shapes of the rotation curves (van den Bosch & Swaters 2000; Navarro & Swaters 2000). Moreover, cuspy density profiles with inner logarithmic slopes as steep as $`1.5`$ also seem to be consistent with the data (van den Bosch & Swaters 2000). A separate concern is not the shape of the inner density profile, but rather the value of the central density. There are indications that CDM halos are too concentrated (Navarro & Swaters 2000; McGaugh et al. 2000; Navarro & Steinmetz 2000; Firmani et al. 2000) in comparison with galactic halos. However, van den Bosch & Swaters (2000) have argued, based on detailed modeling of adiabatic contraction and beam-smearing, that dwarf galaxy concentrations are in fact consistent with the observed distribution in $`\mathrm{\Lambda }`$CDM halos. Thus, although the shape of galactic rotation curves may be not as different from predictions as was thought before, the halo concentrations derived from observations are alarmingly low. The recent observational progress made in this field promises to resolve and clarify this issue in the near future. At larger scales ($`r100`$ kpc), the constraints from weak galaxy-galaxy lensing (Fischer et al. 2000; Smith et al. 2000) should be very useful in constraining the overall profile and concentrations of galactic halos.
In summary, the study presented here is aimed to clarify the issue of convergence of the density profiles of CDM halos. We show that convergence can be reached regardless of the mass resolution, although the convergence scale does depend on the mass resolution: the higher mass resolution results in smaller convergence scale for the same objects, but it does not affect the outer parts of the profile. Our results also indicate that there is a real scatter in shapes of density profiles and halo parameters. Larger systematic studies currently underway may put these conclusions on a firmer footing.
We acknowledge support from the grants NAG- 5- 3842 and NST- 9802787, and also NASA and NSF grants at UCSC. A.V.K. acknowledges support by NASA through Hubble Fellowship grant HF-01121.01-99A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. JSB was supported by NASA LTSA grant NAG5-3525 and NSF grant AST-9802568. JRP acknowledges a Humboldt Award. Computer simulations presented in this paper were done at the National Center for Supercomputing Applications (NCSA), Urbana-Champaign, Illinois and at the Leibniz-Rechenzentrum in Munich.
|
warning/0006/gr-qc0006053.html
|
ar5iv
|
text
|
# Suspensions Thermal Noise in the LIGO Gravitational Wave Detector
## I Introduction
Thermal noise is a fundamental limit to the sensitivity of gravitational wave detectors, such as the ones being built in the use by the LIGO project. Thermal noise is associated with sources of energy dissipation , following the Fluctuation-Dissipation Theorem. Thermal noise comes in at least two important kinds: one due to the brownian motion of the mirrors, associated with the losses in the mirrorsโ material; and another due to the suspension of the mirrors, due to the losses in the wiresโ material. The limits following from these assumptions (losses due to elastic properties of materials) are a lower limit to the noise in the detector, since there may always be other sources of energy dissipation in imperfect clamps, mirror attachments, etc. But the correct calculation of the thermal noise limit is essential to the design of detectors and diagnostics of the already-built detectors. We will deal in this article with thermal noise of suspensions (not of internal modes of the mirrors themselves), and assume only losses due to the elasticity of the suspension wires.
The calculation of thermal noise can be done in several ways ,,,,. All of these follow the Fluctuation-Dissipation Theorem (FDT), but a complication arises because in suspensions there are two sources of energy (gravitational and elastic), but only one of them is โlossyโ (elastic energy). Moreover, the losses in the suspension wires are associated with their bending, and seems to be localized at the top and bottom of the wires. The ways to include these features into the thermal noise calculations are different enough that they have led to some confusion among the gravitational wave community. Also, attention has been paid mostly to the horizontal motion of the suspension, although all modes (angular, transverse, and vertical) appear to some degree into the detectorโs noise. We present a method to calculate thermal noise that allows the prediction of the suspension thermal noise in all its 6 degrees of freedom, from the energy dissipation due to the elasticity of the suspension wires. We also show how the contributions of thermal noise in different directions can be sensed by the interferometer through the laser beam position and direction. The results will follow from the consideration of the coupled equations of the suspension and the continuous wire, first presented in for just the horizontal degree of freedom. We show how this approach encompasses and explains previous ways to approximate the thermal noise limit in gravitational wave detectors. We show how this approach can be extended to more complicated suspensions to be used in future LIGO detectors. To our knowledge, this is the first time the thermal noise of angular degrees of freedom is presented, and that all suspension degrees of freedom are calculated in an unified approach.
Since the full treatment of the problem is somewhat involved, we present first the problem without considering the elasticity of the wire, but adding a second, lossy, energy source to the gravitational energy in the treatment of the mechanical pendulum, and introduce the concepts of โdilution factorsโ, and โeffectiveโ quality factors. We also start with one and two-degrees of freedom suspensions instead of 6-dof. With these tools, most of the issues can be clearly presented and then we follow to the full treatment of the LIGO suspensions, presenting the implications for LIGO.
## II Simple pendulum cases: dilution factors, coupled modes, effective quality factors.
The full treatment of this case, considering the elastic coupling of the wire to the suspension, was presented in . Here, we will present the simpler โmechanicalโ treatment of this case, which will introduce the concepts of โdilution factorsโ, and measured vs. effective quality factors.
### A A simple oscillator with a dissipative energy source
We first recapitulate the calculation of thermal noise in the simplest case, a suspended point mass. The potential energy is $`PE=(1/2)Kx^2`$ and $`F_K=dV/dx=Kx`$. The kinetic energy is $`KE=(1/2)M\dot{x}^2`$. The admittance to an external force $`F_{ext}`$ is given by
$$Yi\omega \frac{x}{F_{ext}}=i\omega \frac{1}{KM\omega ^2}$$
The admittance has a pole at the system eigenfrequency $`w_0^2=K/M`$. If $`K`$ is real, the resonance has an infinite amplitude and zero width. If the spring constant has an imaginary part representing an energy loss, $`KK(1+i\varphi )`$, then the amplitude $`|Y(\omega _0)|`$ is finite, and the peak has a width determined by the complex part of the eigenfrequency $`\omega _0^2=(K/M)(1+i\varphi )`$. The width of the peak is characterized with a quality factor $`Q=1/\varphi `$, and it is usually measured from the free decay time $`\tau `$ of the natural oscillation at the frequency $`\omega _0`$: $`Q=\omega _0\tau /2`$.
The thermal noise is proportional to the real part of the admittance, and thus to $`\varphi `$:
$$\mathrm{}[Y]=\frac{wK\varphi }{(KM\omega ^2)^2+K^2\varphi ^2}$$
We are usually interested in frequencies well above $`\omega _0`$, since the pendulum frequency $`w_0/2\pi `$ in gravitational wave detectors is usually below 1 Hz, and the detectors have their maximum sensitivity at 100 Hz. At those frequencies, the thermal noise is
$$x^2(\omega )=\frac{4k_BT_0}{\omega ^2}\mathrm{}[Y]\frac{4k_BT_0\omega _0^2\varphi }{M\omega ^5}$$
(1)
This how we see that the measured decay of the pendulum mode can be used to predict the suspension thermal noise at gravitational wave frequencies. Some beautiful examples of these difficult measurements and their use for gravitational wave detectors are presented in , for example.
### B A pendulum with two energy sources: the dilution factor.
Next, we consider a suspended point mass, but we now assume there two sources of energy, gravitational and elastic, each with its own spring constant. The potential energy is then $`V=V_g+V_e=(1/2)(K_g+K_e)x^2`$, and
$$\mathrm{}[Y]=\frac{\omega (K_g\varphi _g+K_e\varphi _e)}{((K_g+K_e)M\omega ^2)^2+(K_g+K_e)^2\varphi ^2}$$
If we assume that $`K_gK_e`$, then
$$\mathrm{}[Y]\frac{\omega (K_g\varphi _g+K_e\varphi _e)}{(K_gM\omega ^2)^2+K_g^2\varphi ^2}$$
and at high frequencies
$$x^2(\omega )=\frac{4k_BT_0}{\omega ^2}\mathrm{}[Y]\frac{4k_BT_0(K_g\varphi _g+K_e\varphi _e)}{M^2\omega ^5}$$
If $`\varphi _g=0`$ (โgravity is losslessโ), or at least $`K_g\varphi _gK_e\varphi _e`$, then
$$\frac{4k_BT_0\omega _0^2(K_g/K_e)\varphi _e}{M\omega ^5}$$
where $`\omega _0^2K_g/M`$. We see that is the same expression as if we had just one energy source with a complex spring constant $`K=K_g(1+i(K_e/K_g)\varphi _e)`$. This is why we call the factor $`K_e/K_g`$ the โdilution factorโ: the elastic energy is the one contributing the loss factor to the otherwise loss-free $`K_g`$, but โdilutedโ by the small factor $`K_e/K_g`$. The dilution factor is also equal to the ratio of elastic energy to gravitational energy $`K_e/K_g=V_e/V_g`$. The concept of a dilution factor is very useful because it is usually easier to measure the loss factor $`\varphi _e`$ associated with the elastic spring constant than the quality factor of the pendulum mode. This is because $`K_e`$ is usually a function of the complex Young modulus $`E`$, and the imaginary part of the Young modulus is easily measurable for most fiber materials, and can even be found in tables of material properties. (Of course, there are subtleties to this argument, in particular with thermoelastic or surface losses , but we are assuming the minimum material loss).
### C A point mass suspended from an anelastic wire: calculating the dilution factor
This case is a particular case of the one treated in , and here we just mention it to present the approach taken to the full problem, and present some new relevant aspects.
We want to include the elasticity of the wire in the equations of motion, so we treat the suspension wire as an elastic beam, and then we have pendulum degree of freedom $`x`$, plus the wireโs infinite degrees of freedom $`w(s)`$ of transverse motion. We define a coordinate $`s`$, that starts at the top of the wire $`s=0`$, and ends at the attachment point to the mirror, $`s=L`$. Correspondingly, we will have an eigenfrequency, associated with the pendulum mode, and an infinite series of โviolinโ modes. The potential energy is $`PE=(1/2)_0^L\rho (w^{}(s))^2๐s`$, and the kinetic energy is $`KE=(1/2)_0^L\rho (\dot{w}(s))^2๐s+(1/2)M\dot{x}^2`$. The solutions to the wire equation of motion, with boundary conditions $`w(0)=0`$ and $`w(L)=x`$ are $`w(s)=x\mathrm{sin}(ks)/\mathrm{sin}(kL)`$, and the equation of motion for the mass $`M`$ subject to an external force $`F`$ is $`F=M\omega ^2x+Tw^{}(L)=M\omega ^2x+(T/L)x(kL/\mathrm{tan}(kL))`$. The admittance
$$Y=i\omega \frac{x}{F}=i\omega \frac{T}{L}\frac{\mathrm{tan}kL/kL}{1\omega ^2(ML/T)(\mathrm{tan}kL/kL)}$$
has a pole at the pendulum frequency $`w_p^2T/ML`$, where $`kL1`$, and an infinite number of poles at the violin mode frequencies, at frequencies $`w_n(T/\rho L^2)^{1/2}(1+(\rho L/M)n^2\pi ^2)`$, where $`kL=n\pi +\sqrt{(\rho L/M)}/(n\pi )`$.
The spring โconstantโ associated with the wire and gravityโs restoring force $`K=(T/L)(kL/\mathrm{tan}(kL))`$ is in fact a function of frequency, although it is the usual constant $`T/L`$ for frequencies below the violin modes, where $`kL1`$. At frequencies above the first violin mode, the spring function is not even positive definite, or finite. The function $`K`$ is real at all frequencies because we havenโt added any source of energy loss yet. We introduce energy loss in the system by adding the wire elastic energy to the system, and then assuming a complex Young modulus. The potential energy is now $`PE=(T/2)(_0^Lw^2(s)๐s)+(EI/2)_0^Lw^{\prime \prime 2}(s)๐s`$. The equation of motion for the wire is
$$T\frac{d^2w(s)}{ds^2}EI\frac{d^4w(s)}{ds^4}+\rho \omega ^2w(s)=0,$$
a fourth order equation with boundary conditions $`w(0)=0`$, $`w^{}(0)=0`$, and $`w(L)=x`$. The wire slope at the bottom, $`w^{}(L)`$, is a free parameter (since we are assuming a point mass), and the variation of the Lagrangian with respect to $`w^{}(L)`$ provides the fourth boundary condition for the wire: $`w^{\prime \prime }(L)=0`$. We can find an exact solution for the wire shape as a function of $`x`$, trigonometric functions of $`ks`$, and hyperbolic functions of $`k_es`$, where $`k,k_e`$ are solutions to $`T\kappa ^2EI\kappa ^4+\rho \omega ^2=0`$ which approximate at low frequencies the perfect string wavenumber, $`k^2\rho \omega ^2/T`$ and a constant โelasticโ wavenumber , $`k_e^2T/EI`$ . The distance $`\mathrm{\Delta }=\sqrt{EI/T}`$ is the characteristic elastic distance over which the wire bends, especially at top and bottom clamps. In LIGO test mass suspensions, $`\mathrm{\Delta }`$2mm, a small fraction of $`L=0.45`$m. The approximations $`k^2\rho \omega ^2/T`$ and $`k_e^2T/EI`$ are valid for frequencies that satisfy $`\omega ^2T^2/4EI\rho `$, about 12 kHz for LIGO, so we will use them in the remainder of this article. It is also equivalent to $`k\mathrm{\Delta }1`$.
We also use an approximate solution for the wire shape, good to order $`e^{L/\mathrm{\Delta }}`$($`10^{99}`$ (!) for LIGO):
$`w(s)`$ $`=`$ $`A\left(\mathrm{sin}(ks)k\mathrm{\Delta }\mathrm{cos}(ks)\right)`$ (3)
$`+k\mathrm{\Delta }\left(Ae^{s/\mathrm{\Delta }}+Be^{(Ls)/\mathrm{\Delta }}\right)`$
The coefficients $`A,B`$ are functions of $`x`$ and $`k`$, and thus, functions of frequency:
$`A`$ $`=`$ $`{\displaystyle \frac{x}{\mathrm{sin}(kL)k\mathrm{\Delta }\mathrm{cos}(kL)}}`$
$`B`$ $`=`$ $`xk\mathrm{\Delta }.`$
In the limit $`\mathrm{\Delta }0`$, we recover the perfect wire solution, $`w(s)=x\mathrm{sin}(ks)/sin(kL)`$. The ratio $`B/A`$ measures how much more (or less) the wire bends at the bottom than at the top. The elastic energy is well approximated by the contribution of the exponential terms in the wire shape, at top and bottom: $`PE_e=(1/2)EI(w^{\prime \prime }(s))^2๐s=(1/2)T\mathrm{\Delta }^2(w^{\prime \prime }(s))^2๐s(1/2)Tk^2L(A^2+B^2)`$. At low frequencies where $`kL1`$, the ratio $`B/A\mathrm{\Delta }/L1`$, indicating that the wire bends much more at the top than at the bottom (recall this is a point mass).
The equation of motion for the mass when there is an external force $`F`$ is
$`F`$ $`=`$ $`M\omega ^2x+Tw^{}(L)EIw^{\prime \prime \prime }(L)`$ (4)
$``$ $`M\omega ^2x+Tw^{}(L)`$ (5)
$``$ $`M\omega ^2x+\left({\displaystyle \frac{kT}{\mathrm{tan}kL}}{\displaystyle \frac{1+2k\mathrm{\Delta }\mathrm{tan}kL}{1(k\mathrm{\Delta })/(\mathrm{tan}kL)}}\right)x`$ (6)
The ratio of the elastic force to the gravitational force, $`EIw^{\prime \prime \prime }(L)/Tw^{}(L)`$, is of order $`k\mathrm{\Delta }1`$, and thus it was dropped. If we now consider $`\mathrm{\Delta }`$ complex, then the spring function
$$K=\frac{kT}{\mathrm{tan}kL}\frac{1+2k\mathrm{\Delta }\mathrm{tan}kL}{1(k\mathrm{\Delta })/(\mathrm{tan}kL)}$$
is also complex, and the admittance will have a non-zero real part. At frequencies below the violin modes where $`kL1`$, we have $`K(T/L)(1+\mathrm{\Delta }/L)`$, an expression that suggests a split between a real gravitational spring constant $`K_g=T/L`$ and a complex elastic spring constant $`K_e=T\mathrm{\Delta }/L^2`$. However, this distinction can only be done in the approximation $`\mathrm{\Delta }/L1`$, and low frequencies $`kL1`$. In general, however, we cannot strictly derive separate gravitational and elastic spring constants from their respective potential energy expressions: notice that the total, complex spring constant $`T/(L\mathrm{\Delta })`$ was derived from the variation of the gravitational potential energy term, which becomes complex because we use a wire shape involving the complex distance $`\mathrm{\Delta }`$, satisfying the boundary conditions.
Where the approximations $`\mathrm{\Delta }/L1,kL1`$ are valid, we can consider the case of two separate spring constants and thus a โdilution factorโ for the pendulum loss, $`K_e/K_g=\mathrm{\Delta }/L1/232`$, where the numerical value corresponds to LIGO parameters in Appendix 1. However, if we numerically calculate the exact pendulum mode quality factor, we get $`1/Q_p=461/\varphi `$: this would be the measured Q from a decay time of the pendulum mode, if there are no extra losses. Did we make a โfactor of 2โ mistake? In fact, this factor of 2 has haunted some people in the community (including myself), but there is a simple explanation. Since the elastic complex spring constant is proportional to $`\mathrm{\Delta }`$, and $`\mathrm{\Delta }`$ is proportional to the square root of the Young modulus, then when we make $`E`$ complex $`EE(1+i\varphi )`$, we get $`K_eT\mathrm{\Delta }/L^2(T\mathrm{\Delta }/L^2)(1+i\varphi /2)`$. That is, we get an extra dilution factor of two between the wire loss $`\varphi `$ and the pendulum loss: $`\varphi _p=(K_g/K_e)\mathrm{}K_e/\mathrm{}K_{el}=(K_g/K_{el})\varphi /2=(\mathrm{\Delta }/2L)\varphi \varphi /464`$, very close to the actual value. This teaches us that if the spring constant of the dissipative force is not just proportional to $`E`$, we will get correction factors $`\mathrm{log}K_e/\mathrm{log}E`$.
The thermal noise below the violin modes is well approximated by the thermal noise of a simple oscillator, as in Eqn.1, with natural eigenfrequency $`w_0^2=g/L`$ and loss $`\varphi _0=\mathrm{\Delta }\varphi /(2L)`$. We then call $`\varphi _0=\mathrm{\Delta }\varphi /(2L)`$ the โeffectiveโ loss, in this case equal to the pendulum loss (but we will see this is not always the case).
We saw that the complex spring constant $`T/(\mathrm{\Delta }L)`$was split into gravitational $`T/L`$ and elastic $`T\mathrm{\Delta }/L^2`$ components. However both were derived from the gravitational force $`F_g=Tw^{}(L)`$, since the elastic force, $`F_{el}=EIw^{\prime \prime \prime }(L)`$, was negligible. It is because the wire shape $`w(s)`$ is different due to elasticity, that the function $`Tw^{}(L)`$ is different from the pure gravitational expression $`Tx/L`$. The way we split gravitational and elastic contributions to the spring constant and then got a dilution factor, is only valid at low frequencies. So the argument we posed in the previous section about a dilution factor applied to the calculation in the thermal noise in the gravitational wave band is in priciple not applicable here, especially when taking into account that the total force was contributed by the variation of just the gravitational potential energy, with the elasticity in the wire shape. However, using the wire shape without low frequency approximations, we can numerically evaluate the integrals that make up the potential and elastic energies (using a real $`\mathrm{\Delta }`$), and compare the ratio $`V_{el}/V_g=T(w^{})^2๐s/EI(w^{\prime \prime })^2๐s`$ with the โlow frequencyโ dilution factor $`\mathrm{\Delta }/L`$. We show the calculation of elastic and gravitational potential energies, and their ratio, in Fig.1. At low frequencies, the ratio is constant, and equal to $`\mathrm{\Delta }/2L`$: this is the dilution factor between the wire loss $`\varphi `$ and the pendulum loss $`\varphi _p`$, also the one to use for a simple-oscillator approximation of the thermal noise. It is not the ratio of the โgravitationalโ and โelasticโ spring constants at low frequencies, but as we explained, we had no reason to expect that, since $`V_e/V_gK_e/K_g`$. At higher frequencies, the ratio $`V_{el}/V_g`$ is not constant, and it gives correctly the dilution factors for the quality factors of the violin modes. Notice that the loss at the violin modes increases with mode number, as noted in : $`\varphi _n=(\mathrm{\Delta }/L)(1+n^2\pi ^2\mathrm{\Delta }/L)=(\mathrm{\Delta }/L)(1+\rho L\mathrm{\Delta }\omega _n^2/T)`$, and this anharmonic behavior is well followed by the energy ratio.
In summary, the concept of dilution factor is strictly true only when the total restoring force can be split into two forces, one lossless and one dissipative, both represented with spring constants. In the general case, if we can only split the potential energy into two terms, one lossless and another dissipative, then the ratio of the energies calculated as a function of driving frequency is the exact dilution โfactorโ. Moreover, this ratio can be calculated as a function of frequency, and then we get the different dilution factors for all the modes in the system. This is an important lesson that also we will use more extensively in suspensions with more coupled degrees of freedom.
### D A suspended extended mass: coupled degrees of freedom and observed thermal noise
We now consider an extended mass instead of a point mass, with a single generic dissipative energy source. The pendulum motion is described with the horizontal displacement of its center of mass $`x`$, and the pitch angle of the mass, $`\theta `$, as in the side view of a LIGO test mass, in Fig.7. The kinetic energy is $`KE=(1/2)(M\dot{x}^2+J\dot{\theta }^2)`$. Instead of a spring constant, we have a 2x2 spring matrix. The potential energy is
$`PE`$ $`=`$ $`{\displaystyle \frac{T}{2}}(L\alpha ^2+h\theta ^2)`$ (7)
$`=`$ $`{\displaystyle \frac{1}{2}}(K_{xx}x^2+2K_{x\theta }x\theta +K_{\theta \theta }\theta ^2)`$ (8)
where $`\alpha =(x+h\theta )/L`$ and $`\theta `$ are the normal coordinates. The point $`x+h\theta `$ is the point where the wire is attached to the mirror, and the angle $`\alpha `$ is the angle the wire makes with the vertical. If we only consider gravitational forces, $`K_{xx}=T/L,K_{x\theta }=Th/L`$ and $`K_{\theta \theta }=Th(L+h)/L`$. However, we will assume that the elements of the spring matrix can be complex, and each has its own different imaginary part. The eigenfrequencies are the solutions to the equation $`(K_{xx}M\omega ^2)(K_{\theta \theta }J\omega ^2)K_{x\theta }^2=0`$, or
$`2MJ\omega _\pm ^2=`$ $`(MK_{\theta \theta }+JK_{xx})`$ (10)
$`\pm \left((MK_{\theta \theta }JK_{xx})^2+4MJK_{x\theta }^2\right)^{1/2}`$
In order to calculate $`x^2(\omega )`$ (the Brownian motion of the center of mass), we need to calculate the admittance $`x/F`$ to a horizontal force applied at the center of mass. In order to calculate $`\theta ^2(\omega )`$, we need the admittance $`\theta /N`$ to a torque applied around the pitch axis. If the spring constants are complex, then the admittances are complex and we can calculate their real parts, and the thermal noise determined by them:
$$x^2(\omega )=\frac{4k_BT_0}{\omega ^2}\mathrm{}\left[i\omega \frac{1J\omega ^2/K_{\theta \theta }}{K_{xx}(1\omega ^2/\omega _+^2)(1\omega ^2/\omega _{}^2)}\right]$$
$$\theta ^2(\omega )=\frac{4k_BT_0}{\omega ^2}\mathrm{}\left[i\omega \frac{1M\omega ^2/K_{xx}}{K_{\theta \theta }(1\omega ^2/\omega _+^2)(1\omega ^2/\omega _{}^2)}\right]$$
where the eigenfrequencies are now complex: $`\omega _\pm ^2\omega _\pm ^2(1+i\varphi _\pm )`$. The quality factors measurable from the free decay of each of the eigenfrequencies are $`Q_\pm =\omega _\pm \tau _\pm /2`$.
At frequencies larger than any of the eigenfrequencies, we obtain
$`x^2(\omega )`$ $``$ $`{\displaystyle \frac{4k_BT_0}{K_{xx}\omega ^5}}((\omega _+^2\varphi _++\omega _{}^2\varphi _{})\varphi _{\theta \theta }(K_{\theta \theta }/J))`$ (11)
$`\theta ^2(\omega )`$ $``$ $`{\displaystyle \frac{4k_BT_0}{K_{\theta \theta }\omega ^5}}((\omega _+^2\varphi _++\omega _{}^2\varphi _{})\varphi _{xx}(K_{xx}/M))`$ (12)
The system may have the two eigenfrequencies close in value if $`h(L+h)J/M`$ (see Fig.2), but for $`hJ/ML`$, we have $`\omega _{}^2K_{\theta \theta }/J=Th/J`$ and $`\omega _+^2K_{xx}/ML^2=T/ML`$; and for $`h(L+h)J/M`$, $`\omega _+^2K_{\theta \theta }/J=Th/J`$ and $`\omega _{}^2K_{xx}/ML^2=T/ML`$. In both limits, two terms cancel in the sum of loss factors in the formulas above ($`\omega _{}^2\varphi _{}\varphi _{\theta \theta }K_{\theta \theta }/J`$ for small $`h`$, for example) and we see that
$$x^2(\omega )\frac{4k_BT_0\omega _p^2\varphi _{xx}}{Mw^5}\mathrm{and}\theta ^2(\omega )\frac{4k_BT_0\omega _\theta ^2\varphi _{\theta \theta }}{J\omega ^5}$$
Thus, even though it is a coupled system, thermal noise in $`x`$ is always associated mostly to $`K_{xx}(1+i\varphi _{xx})`$ and the pendulum eigenfrequency $`\omega _p^2=K_{xx}/MT/ML`$; and thermal noise in $`\theta `$ is always associated with $`K_{\theta \theta }(1+i\varphi _{\theta \theta })`$ and the pitch eigenfrequency $`\omega _\theta ^2=K_{\theta \theta }/JTh/J`$. For both degrees of freedom $`x`$ and $`\theta `$, we obtain the thermal noise of single-dof systems. Unfortunately, neither limit (small or large $`h`$ with respect to $`J/ML`$) applies to the suspension parameters in LIGO test masses, and, more importantly, even though the approximation for the eigenfrequencies is relatively good for most values of $`h`$, the approximation we used for the losses is not (Fig.2). The measurable quality factors give us $`\varphi _\pm `$, but we need $`\varphi _{xx}`$ and $`\varphi _{\theta \theta }`$ to use in the thermal noise of the pendulum, and these cannot be precisely calculated from $`\varphi _\pm `$ unless we know $`\varphi _{x\theta }`$, or a way to relate it to the other loss factors. We will do this in the next section, using the elasticity of the wire.
Notice that the forces and torques we have used to calculate the admittances $`Y_x`$ and $`Y_\theta `$, will each produce both displacement and rotation of the pendulum. This means that the thermal noise in displacement and angle are not uncorrelated. This can be exploited to find a point other than the center of mass where the laser beam in the interferometer would be sensing less displacement thermal noise than at the center of the mirror, as was done following a somewhat different logic in . If we were to calculate the thermal noise at a point a distance $`d`$ above the center of mass, we then need to calculate the admittance of the velocity of that point ($`i\omega (x+d\theta )=i\omega \chi `$) to a horizontal force applied at that point. The equations of motion are
$`F=`$ $`(K_{xx}M\omega ^2)x+K_{x\theta }\theta `$
$`Fd=`$ $`K_{x\theta }x+(K_{\theta \theta }J_y\omega ^2)\theta `$
and then the thermal noise is
$`\chi ^2(\omega )`$ $`=`$ $`{\displaystyle \frac{4k_BT_0}{\omega ^2}}\mathrm{}\left[i\omega {\displaystyle \frac{(x+d\theta )}{F}}\right]`$ (13)
$`=`$ $`{\displaystyle \frac{4k_BT_0}{\omega ^2}}\mathrm{}\left[Y_{xx}+d^2Y_{\theta \theta }+2dY_{x\theta }\right]`$ (14)
$`=`$ $`x^2(\omega )+d^2\theta ^2(\omega )+2d{\displaystyle \frac{4k_BT_0}{\omega ^2}}\mathrm{}(Y_{x\theta })`$ (15)
where $`Y_{xx}`$ is the admittance of $`x`$ to a pure force $`F`$, $`Y_{\theta \theta }`$ is the admittance of $`\theta `$ to a pure torque $`N`$, and $`Y_{x\theta }`$ is the admittance of a displacement $`x`$ to a pure torque $`N`$, equal to the admittance of $`\theta `$ to a pure force $`F`$. There is an optimal distance $`d`$ below the center of mass for which the thermal noise $`\chi ^2(\omega )`$ is a minimum: this distance is $`d=\mathrm{}(Y_{x\theta })/\mathrm{}(Y_{\theta \theta })`$. The resulting thermal noise is
$$\chi _{\mathrm{min}}^2(\omega )=x^2(\omega )\frac{4k_BT_0}{\omega ^2}\frac{(\mathrm{}Y_{x\theta })^2}{\mathrm{}Y_{\theta \theta }}$$
which is less than the thermal noise $`x^2(\omega )`$ observed at the center of mass. However, the expression obtained for the distance $`d`$ is frequency-dependent: that means we have to choose a frequency at which to optimize the sampling point.
Summarizing, we have shown that whenever there are coupled motions, the thermal noise sensed at a point whose position depends on both coordinates is not the sum in quadrature of the two thermal noise ($`x^2`$ and $`d^2\theta ^2(\omega )`$ in our case), but a combination that depends on the โcross-admittanceโ. Moreover, the thermal noise of each degree of freedom cannot in general be calculated just from the measured quality factors if the modes are coupled to each other strongly enough.
### E A 2-DOF pendulum suspended from a continuum wire
We add to the previous 2-DOF pendulum a continuum wire, to be able to add the losses due to the wireโs elasticity, and calculate modal and effective quality factors, as well as the point on the mirror at which we can sense the minimum thermal noise.
If we add elasticity to the problem, as in , the potential energy is $`PE=(T/2)(_0^Lw^2(s)๐s+h\theta ^2)+(EI/2)_0^Lw^{\prime \prime 2}(s)๐s`$. The boundary conditions for the wire equation are $`w(0)=0,w(L)=x+h\theta `$, and $`w^{}(0)=0,w^{}(L)=\theta `$. The equations of motion for the pendulum are
$`M\omega ^2x+Tw^{}(L)EIw^{\prime \prime \prime }(L)=`$ $`F`$ (16)
$`J\omega ^2\theta +EI(w^{\prime \prime }(L)+hw^{\prime \prime \prime }(L))=`$ $`N`$ (17)
In order to complete the equations of motion of the pendulum, we need the shape of the wire at the bottom end. For this, we use the shape given by the expression in Eqn.3, but this time the top and bottom weights are given by
$$A=\frac{x+(h+\mathrm{\Delta })\theta }{D}$$
$`B={\displaystyle \frac{1}{kD}}(`$ $`kx(\mathrm{cos}(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))`$
$`+\theta (\mathrm{sin}(kL)+k(h\mathrm{\Delta })\mathrm{cos}(kL))`$
with $`D=\mathrm{sin}(kL)2k\mathrm{\Delta }\mathrm{cos}(kL)`$ and $`\mathrm{\Delta }=\sqrt{EI/T}`$ as we used earlier. With the shape known, we can write the equations for $`x,\theta `$ with a spring matrix:
$`(K_{xx}M\omega ^2)x+K_{x\theta }\theta =`$ $`F`$
$`K_{\theta x}x+(K_{\theta \theta }J_y\omega ^2)\theta =`$ $`N`$
where the spring functions are
$`K_{xx}`$ $`=`$ $`Tk(\mathrm{cos}(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))/D`$ (18)
$`K_{\theta \theta }`$ $`=`$ $`T(h+\mathrm{\Delta })(\mathrm{sin}(kL)+k(h\mathrm{\Delta })\mathrm{cos}(kL))/D`$ (19)
$`K_{x\theta }`$ $`=`$ $`K_{\theta x}=Tk(h+\mathrm{\Delta })(\mathrm{cos}(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))/D`$ (20)
At this point, even though the expressions are complicated, we can calculate the complex admittances $`Y_{xx}=i\omega x/F,Y_{\theta \theta }=i\omega \theta /N`$ using a complex $`E`$ and $`\mathrm{\Delta }`$. The analytical expressions for the admittances are quite involved, but we can always calculate numerically the thermal noise associated with any set of parameters. We can also calculate the widths of the peaks in the admittance, which would correspond to measurable quality factors for the pendulum, pitch, and violin modes. The plots presented in Figs. 2 and 3 were calculated using these solutions.
Fig.2 shows that the frequency and quality factor of the pendulum and pitch modes vary significantly with the pitch distance $`h`$. At frequencies close to the pendulum eigenfrequencies, the thermal noise spectral densities show peaks at both frequencies. However, at higher frequencies, the thermal noise $`x^2(\omega )`$ can always be approximated by the thermal noise of a simple oscilaltor as in Eqn.1, with an โeffectiveโ quality factor that fits the amplitude at high frequencies to the position of the single peak. We can similarly define an effective quality factor for the thermal noise in $`\theta ^2(\omega )`$. We show in Fig.3 the actual and approximated thermal noise, with their corresponding effective quality factors found to fit best at 50 Hz. The effective quality factor at 50 Hz can be calculated as a function of pitch distance, and we show this calculation in Fig. 2. The effective pitch quality factor is well approximated, for any pitch distance, by the measurable pitch quality factor, while the pendulum effective quality factor is close to the measurable quality factor of the pendulum mode only at very small, or very large pitch distances. For the LIGO pitch distance of 8mm, the measurable pendulum quality factor is 10 times lower than the effective quality factor, and would then give a pessimistic estimate of thermal noise amplitude.
Low frequency approximation. At low frequencies, where $`kL1`$, we have expressions that can help us understand how the elasticity loss factor contributes to the effective quality factors, as well as to the pendulum and pitch modes. We trade this gain in simplicity for the loss of expressions valid at or above violin mode resonances.
The low frequency limit of the spring constants in Eqns. 20 is
$`K_{xx}`$ $`=`$ $`T/(L2\mathrm{\Delta })`$
$`K_{\theta \theta }`$ $`=`$ $`T(h+\mathrm{\Delta })(L+h\mathrm{\Delta })/(L2\mathrm{\Delta })`$
$`K_{\theta x}`$ $`=`$ $`K_{x\theta }=T(h+\mathrm{\Delta })/(L2\mathrm{\Delta }).`$
If we assume $`\mathrm{\Delta }`$ has an imaginary part related to the material $`\varphi `$: $`\mathrm{\Delta }\mathrm{\Delta }(1+i\varphi /2)`$, then we get complex spring constants. If we use these complex spring constants in Eq.10 we can calculate the loss factors of the pendulum and pitch mode, $`\varphi _p`$ and $`\varphi _\theta `$; and if we use them in Eqns. 12 we can get the effective quality factors. Using $`h/L1`$ and $`\mathrm{\Delta }/L1`$, we get $`Q_{x\mathrm{eff}}=1/\varphi _{xx}\mathrm{\Delta }\varphi /L`$ and $`Q_{\theta \mathrm{eff}}=1/\varphi _{\theta \theta }\mathrm{\Delta }\varphi /2(h+\mathrm{\Delta })`$. This represents a โdilution factorโ in displacement of $`\mathrm{\Delta }/L`$ and in pitch of $`\mathrm{\Delta }/2(h+\mathrm{\Delta })`$. These approximations fit very well the values shown in Fig. 2.
At low frequencies, the equations of motion for $`x,\theta `$ can be derived from a potential energy
$$PE_{kL1}=\frac{T}{2}\left(\frac{(x+(h+\mathrm{\Delta })\theta )^2}{L2\mathrm{\Delta }}+(h+\mathrm{\Delta })\theta ^2\right).$$
Using a complex $`\mathrm{\Delta }`$, this gives us the complex spring constants that we can use to get mode and effective quality factors. We would like to break up this potential energy into a gravitational part and an elastic part, corresponding to a real gravitational spring constant (independent of $`\mathrm{\Delta }`$) and a lossy elastic spring constant. We know that if we take the limit $`\mathrm{\Delta }0`$ in $`PE_{kL1}`$, we obtain the regular potential energy $`PE_g`$ in Eq.8 for a 2 DOF pendulum without elasticity. Thus, we are tempted to say that the elastic energy is the remainder, proportional to $`\mathrm{\Delta }`$, and thus having a complex spring constant when we consider a complex $`\mathrm{\Delta }`$. According to this argument, we get $`K_{gxx}=T/L`$ and $`K_{exx}=2T\mathrm{\Delta }/L^2`$. This would then give us a dilution factor for the displacement loss $`K_e/K_g2\mathrm{\Delta }/L`$. This is the factor by which the imaginary part of $`K_{el}`$ is diluted; however, as explained before, we pick up another factor of two due to $`K_e`$ being proportional to $`\sqrt{E}`$. Thus, the dilution factor between the effective quality factor and the wire quality factor is $`\mathrm{\Delta }/L`$.
Energy Ratios and the Dilution Factor.We have seen that there is another way of identifying the โgravitationalโ and โelasticโ terms in the potential energy, using the actual potential energy expressions from which the equations of motion were derived: $`PE_g=(T/2)(_0^Lw^2(s)๐s+h\theta ^2)`$ and $`PE_{el}=(EI/2)_0^Lw^{\prime \prime 2}(s)๐s`$. For any given applied force, or torque, we can solve the equations of motion for $`x,\theta `$ and $`w(s)`$. (We donโt need to invoke a low frequency approximation to do this calculation.) Then, if we calculate the ratio of elastic to gravitational potential energy for a unit applied force, we get a function which is frequency dependent, and is equal to the dilution factors for the pitch mode at the pitch frequency, for the pendulum mode at the pendulum frequency, and for the effective quality factor at high frequencies, $`\mathrm{\Delta }/L`$. We show the energy values and ratios for different values of the pitch distance in Fig. 4.
Potential Energy Densities.There is another interesting calculation we can do with the solution obtained for the wire shape, and that is to find out where in the wire the elastic potential energy is concentrated. In other words, we want to find a relationship between the variation of the dilution factor with frequency, and the curvature of the wire, mostly at the top and bottom clamps. Since both the gravitational energy and the elastic energy involve integrals over the wire length, we can define energy densities along the wire, and calculate a cumulative integral from top to bottom. The gravitational potential energy has also a term $`Th\theta ^2/2`$: we define a ratio $`R=Th\theta ^2/(_0^Lw^2๐s)`$, indicating the relative contribution of this โpitchโ term. From Fig.5, we observe that the gravitational potential energy density is distributed quite homogenously along the wire, even at the first violin mode. However, the pitch term, which can be considered a โbottomโ contribution, contributes most of the gravitational energy when the system is excited at the pitch eigenfrequency, but also in several other cases at the pendulum frequency and at low frequencies. The elastic energy density is concentrated at top and bottom portions of length $`2\mathrm{\Delta }`$. At low frequencies, the top contributes the most; at the pendulum eigenfrequency, the relative contributions depend strongly on $`h`$, but the bottom contributes at least half the energy; at the pitch eigenfrequency, the bottom contributes more than 99% of the energy; at higher frequencies, including the violin modes, top and bottom contribute equally.
Motion of points away from center of mass. We discussed previously how it was possible to find a point whose thermal noise displacement was smaller than the thermal noise displacement of the center of mass. Now that we have expressions for the complex spring functions, we can find the optimal point and discuss the differences. The cross admittance in Eq.15 is
$`Y_{x\theta }`$ $`=`$ $`i\omega {\displaystyle \frac{x}{N}}=i\omega {\displaystyle \frac{\theta }{F}}`$
$`=`$ $`{\displaystyle \frac{K_{x\theta }}{(K_{xx}M\omega ^2)(K_{\theta \theta }J\omega ^2)}}`$
$``$ $`{\displaystyle \frac{K_{x\theta }}{MJ\omega ^4}}`$
and then the optimal point (otimized at frequencies in the gravitational wave band, above pendulum modes) is $`d_0=\mathrm{}(Y_{x\theta })/\mathrm{}(Y_{\theta \theta })J/ML`$. Notice that even though the optimal distance was deduced from the thermal noise expressions, which all involve loss factors, the optimal distance only depends on mechanical parameters. As first explained in , the interpretation of this distance is that when the pendulum is pushed at that point by a horizontal force, the wire doesnโt bend at the bottom clamp, producing less losses. The fact that we can recover the result from the FDT is another manifestation of the deep relationship between thermal fluctuations and energy dissipation. We show in Fig6 the dependence of the thermal noise at 50 Hz on the point probed by the laser beam on the mirror, and the ratio of the thermal noise for $`d=0`$ and $`d=d_0`$ at all frequencies. As expected, since the integrated rms has to be the the same for any distance at which we sense the motion, the fact that the spectral density is smaller at 50 Hz if $`d=d_0`$ means that the noise will be increased at some other frequencies: this happens mainly at frequencies below the pendulum modes.
There are many lessons to be learned from this exercise, but perhaps the most important one is that the explicit solutions to the equations of motion have many different important results:
* using the solutions to calculate the elastic and gravitational potential energies allows us to calculate a โdilution functionโ of frequency, equal to the dilution factor at each of the resonant modes of the system, as well as to the most important effective dilution factor at frequencies in the gravitational wave band;
* we can calculate energy densities along the wire to identify the portions of the wire most responsible for the energy loss and thus the thermal noise;
* we can use low frequency approximations to find out expressions for dilution factors that can be found using other methods, explaining in this way subtleties like factors of two;
* we can calculate the admittance of an arbitrary point in the mirror surface to the driving force, and thus find out improvements or degradation of observed noise due to beam misalignments.
## III The LIGO suspensions: thermal noise of all degrees of freedom
We will now calculate the solutions to the equations of motion for the six degrees of freedom of a LIGO suspended test mass, and then use the solutions to calculate the thermal nosie of all degrees of freedom, as well as the observed tehrmal noise in the gravitational wave detector.
The mirrors at LIGO are suspended by a single wire looping around the cylindrical mass, attached at the top at a distance smaller than the mirror diameter, to provide a low yaw eigenfrequency. This is equivalent to having a mass suspended by two wires, attached slightly above the horizontal plane where the center of mass is. The mirrorโs 6 degrees of freedom are the longitudinal and transverse horizontal $`x`$ and $`y`$, and the vertical $`z`$, displacements of the center of mass; the pitch $`\theta `$ and yaw $`\varphi `$ rotations around the $`y`$ and $`z`$ axis, respectively, and the roll $`\psi `$ around the longitudinal $`x`$ axis. We show the coordinate system used and the relevant dimensions in Fig.7. The parameters used in the calculation presented are those for LIGO test mass suspensions (Large Optics Suspensions). The mass of the cylindrical mirror is 10.3 Kg, the diameter is 25cm, and the thickness 10cm. The cylindrical wires are made of steel with density $`\rho =7.8\times 10^3\mathrm{kg}/\mathrm{m}^3`$ and 0.62mm diameter. We assumed a complex Young modulus $`E=2.1\times 10^11(1+10^3i)\mathrm{kg}/\mathrm{m}^2`$. The vertical distance between the center of mass and the top clamps is $`l=45`$cm, the wires are attached to the mass a distance $`h=8.2`$mm above the center of mass. The distance between the top attachment points is $`2a=33.3`$mm. (In the previous examples where one wire was used, we assumed the same wire material and the same test mirror, but we used a $`0.88`$mm radius, so the stress in the wires remained constant.)
Each wire element has displacement in a 2-dimensional plane transverse to the wire, $`\stackrel{}{w_i}_{}(s)`$ and a longitudinal displacement along the wire, $`w_{i}^{}{}_{}{}^{}(s)`$. The kinetic energy is given by
$`KE={\displaystyle \frac{1}{2}}(`$ $`{\displaystyle \underset{i=1,2}{}}{\displaystyle _0^L}\rho |\dot{\stackrel{}{w}_i}(s)|^2๐s`$
$`+M(\dot{x}^2+\dot{y}^2+\dot{z}^2)+J_x\dot{\psi }^2+J_y\dot{\theta }^2+J_z\dot{\varphi }^2)`$
The potential energy is given by the sum of the axial strain energy and the bending (transverse) strain energy in each wire:
$`PE_i=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _0^L}๐s\left(Tw_{i}^{}{}_{}{}^{}{}_{}{}^{2}+EA(w_{i}^{}{}_{}{}^{})^2\right)`$
$`+{\displaystyle \frac{1}{2}}{\displaystyle _0^L}๐sEI\left(w_{i}^{\prime \prime }{}_{}{}^{}\right)^2`$
plus the energy involved in rotating the mass:
$$PE_M=T(h\mathrm{cos}\alpha (\theta ^2+\psi ^2)b\mathrm{sin}\alpha (\varphi ^2+\psi ^2)).$$
The wires will be attached at the top ($`s=0`$) at the coordinates $`\stackrel{}{w}_i=(0,\pm a,l)`$, where $`l`$ is the vertical distance of the top support from the equilibrium position of the center of mass. The wiresโ transverse slopes at the top will be zero. At the bottom the wires are clamped to the mass a distance $`h`$ above the center of mass, and a distance $`2b`$ on the y-direction between the wires on each side of the mass. The angle $`\alpha =\mathrm{arctan}((ba)/(lh))`$ is the angle at which the wires are slanted from top to bottom when looking at the mass along the optical axis. If $`b=a`$, the wires hang vertically. The length of the wires is $`L^2=(ba)^2+(lh)^2`$. The tension in each wire is $`T=Mg\mathrm{cos}\alpha /2`$. The position of the bottom attachments when the mirror is moving with a motion described by $`(x,y,z,\theta ,\varphi ,\psi )`$ are $`w_i(L)=(x+h\theta \pm b\varphi ,yh\psi ,z\pm b\psi )`$, and the slopes at the bottom are $`w_i^{}(L)=(\theta ,\psi ,0)`$.
If we express the wire transverse and longitudinal displacements in the $`x,y,z`$ coordinate system, we have $`w_{i}^{}{}_{}{}^{}=w_z\mathrm{cos}\alpha \pm w_y\mathrm{sin}\alpha `$, and $`w_{i}^{}{}_{}{}^{}=\sqrt{w_{i}^{}{}_{x}{}^{2}+(w_{i}^{}{}_{y}{}^{}\mathrm{cos}\alpha \pm w_{i}^{}{}_{z}{}^{}\mathrm{sin}\alpha )^2}`$ and the equations of motion become non-linear. In order to keep the problem simple, without losing any degree of freedom, we will then consider two different cases: (i) the wire only has displacements in the $`x`$ direction, and the mirror moves in $`x,\theta ,\varphi `$ degrees of freedom; and (ii) the wire only has displacements in the $`y,z`$ directions, and the mirror moves in $`y,z,\psi `$ degrees of freedom. We analyze these cases separately.
### A Longitudinal, Pitch and Yaw Thermal Noise
The boundary conditions for the wires at the top are zero displacements and slopes, and at the bottom attachment to the mass, $`w_{}^{}{}_{i}{}^{}(L)=x+h\theta \pm \varphi ,w_{}^{}{}_{i}{}^{}(L)=\theta `$.
We combine the wiresโ transverse displacements into $`w_\pm (s)=(w_1(s)\pm w_2(s))/2`$, then the boundary conditions at the bottom are $`w_+(L)=xh\theta ,w_+^{}(L)=\theta ,w_{}(L)=b\varphi ,w_{}^{}(L)=0`$. The solutions to the wire equations of motion and the boundary conditions are (up to order $`e^{L/\mathrm{\Delta }}`$) will then be $`w_\pm (s)`$ as in Eqn.3, with
$`A_{}`$ $`=`$ $`\varphi b/D`$
$`B_{}`$ $`=`$ $`\varphi b(\mathrm{cos}(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))/D`$
$`A_+`$ $`=`$ $`(x+(h+\mathrm{\Delta })\theta )/D`$
$`B_+`$ $`=`$ $`\mathrm{\Delta }(kx(\mathrm{cos}(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))`$
$`+\theta (\mathrm{sin}(kL)+k(h\mathrm{\Delta })\mathrm{cos}(kL)))/D`$
where $`D=\mathrm{sin}(kL)2k\mathrm{\Delta }\mathrm{cos}(kL)`$. The equations for the mass transverse dof subject to a force $`F_x`$ and torques $`N_y,N_z`$, are
$`Fx`$ $`=`$ $`M\omega ^2x2EIw_+^{\prime \prime \prime }(L)+2T\theta `$ (21)
$`N_y`$ $`=`$ $`J_y\omega ^2\theta +2EI\left(w_+^{\prime \prime }(L)+hw_+^{\prime \prime \prime }(L)\right)`$ (23)
$`+2Th\theta (1\mathrm{cos}\alpha )`$
$`N_z`$ $`=`$ $`J_z\omega ^2\varphi 2EIbw_{}^{\prime \prime \prime }(L)2Tb\varphi \mathrm{sin}\alpha `$ (24)
We see that the combination $`w_+(s)`$ is associated with the $`x,\theta `$ degrees of freedom just as for the single wire case, while the combination $`w_{}(s)`$ is associated with the yaw degree of freedom $`\varphi `$. This is easily understood when imagining the wires moving back and forth โin phaseโ ($`w_{}=0`$), producing displacement and pitch but nt yaw; while if they move back and forth in opposition ($`w_+=0`$), then the only effect is into mirrorโs yaw. Thus, we can solve the equations for $`w_{}(s)`$ and $`\varphi `$ separately from the equations for $`\omega _+(s),x`$ and $`\theta `$: we will do so in the next parapgraphs.
Yaw angular thermal noise. The admittance of yaw $`\varphi `$ to a torque $`N_z`$ is
$$Y_\varphi =i\omega \frac{1}{K_\varphi 2Tb\mathrm{sin}\alpha J_z\omega ^2}$$
with $`K_\varphi =(2Tb^2/L)(kL)(\mathrm{cos}(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))/(sin(kL)2k\mathrm{\Delta }\mathrm{cos}(kL))`$. As usual, when $`\mathrm{\Delta }=\sqrt{EI/T}`$ is complex, the admittance is complex and using the fluctuation-dissipation theorem we obtain $`\varphi ^2(f)=(4k_BT_0)\mathrm{}(Y_\varphi )/\omega ^2`$.
The effect of the tilted wires with $`\alpha =(ba)/L>0`$ is to lower the restoring force, and thus the resonance frequency: in LIGO suspensions, the frequency is 0.48 Hz instead of 1.32 Hz if $`b=a`$. However, since $`\mathrm{sin}\alpha 0`$ decreases the gravitational restoring force but not the elastic force, the dilution factor $`K_e/K_g`$ increases, and so does the thermal noise. At frequencies where $`kL1`$, $`K_\varphi (2Tb^2/L)(1+2\mathrm{\Delta }/L)`$. The thermal noise at frequencies below the violin modes, where $`kL1`$, is well approximated by the thermal noise of a single oscillator with resonance frequency $`w_\varphi ^2=2Tab/JL`$ and quality factor $`Q_\varphi =(L/\mathrm{\Delta })(a/b)(1/\varphi _w)`$, where $`EE(1+i\varphi _w)`$. Thus, the โdilution factorโ is $`(\mathrm{\Delta }/L)(b/a)`$ (=1/72 for LIGO parameters, where $`b/a=7.5`$). As in the case of a simple pendulum, this dilution factor is half of the ratio of the elastic spring constant $`K_e=(2Tb^2/L)(2\mathrm{\Delta }/L)`$ to the gravitational spring constant $`K_g=2Tab/L`$, because the elastic spring constant has an โextraโ dilution factor of 2: if $`EE(1+i\varphi _w)`$, then $`\mathrm{\Delta }\mathrm{\Delta }(1+i\varphi _w/2)`$ and $`K_eK_e(1+i\varphi _w/2)`$. However, the ratio of elastic potential energy $`_0^LEI(w_{}^{\prime \prime })^2๐s`$ to gravitational potential energy $`V_g=_0^LT(w_{}^{})^2๐sTb\mathrm{sin}\alpha \varphi ^2`$ is equal to the โrightโ dilution factor $`(b/a)(\mathrm{\Delta }/L)1/44`$ at low frequencies, as shown in Fig. 8. The energy ratio also gives us the right dilution factor at the violin frequencies ($`2\mathrm{\Delta }/L`$).
The yaw angular thermal noise may be seen in the detectorsโ gravitational wave signal if the beam hits a mirror at distance $`d`$ to either side of the center of mass, or if it hits the mirror in a direction an angle $`\gamma `$ away from longitudinal. Considering both cases, the sensed thermal noise will be given by $`\chi ^2(\omega )=(d\mathrm{cos}\gamma +H\mathrm{sin}\gamma /2)^2\varphi ^2(\omega )`$, where $`H`$ is the thickness of the mirror:
$`\chi ^2(\omega )`$ $`=`$ $`(d\mathrm{cos}\gamma +(H/2)\mathrm{sin}\gamma )^2\varphi ^2(\omega )`$
$``$ $`{\displaystyle \frac{4k_BT}{\omega ^5}}(d\mathrm{cos}\gamma +(H/2)\mathrm{sin}\gamma )^2{\displaystyle \frac{2Tb^2}{J_xL}}{\displaystyle \frac{\mathrm{\Delta }}{L}}\varphi `$
where the approximation is valid between the pendulum mode and the first violin mode, 1Hz-50Hz. At 160 Hz, where the maximum sensitivity of $`\sqrt{h^2(f)}=2.5\times 10^{23}/\sqrt{\mathrm{Hz}}`$ is expected, the yaw thermal noise is $`\sqrt{\varphi ^2}(160\mathrm{H}\mathrm{z})=2.9\times 10^{19}\mathrm{rad}/\sqrt{\mathrm{Hz}}`$. If the yaw thermal nose is to be kept an order of magnitude below the dominant noise source, then it is required that $`d`$1cm and $`\gamma 10^o`$.
Notice that the mirror will always be aligned normal to the laser beam to make the optical cavities resonant; however, what matters is the beam direction with respect to the coordinate system defined by the local vertical and the plane defined by the mirror in equilibrium. Presumably there will be forces applied to align the mirror, but in principle they have no effect on the response of the mirror to an oscullatory driving force such as the one we imagine in the beamโs direction, to calculate the admittance. Thus the requirement on $`\gamma 10^o`$ is on the position of the mirror when there are no bias forces acting, with respect to the ultimate direction of the beam. The beamโs direction must be within 1$`\mu `$rad of the normal to the aligned mirror to keep the beam aligned on mirrors 4km apart, but that doesnโt mean that the mirror has not been biased by less than $`10^o`$ to get it to the final position.
Pitch and displacement thermal noise. We now solve the equations for $`w_+(s),\theta `$ and $`x`$. The equations for these degrees of freedom in Eqns. 24 are exactly the same as for the pendulum suspended on a single wire (Eqn. 17), except for the addition of a softening term to the torque equation, due to the tilted wires; and factors of two due to the two wires (with about half the tension) instead of a single wire. The extra term in the torque equation is a negligible contribution to the real part of $`K_{\theta \theta }`$, at the level of 1% for LIGO parameters. Therefore, the conclusions we obtained, with respect to the optimization of the beam location on the mirror, and the difference between effective and measurable quality factors, are equally valid here. The spring constants we obtained in Eqns. 20 involve now a factor $`2T=Mg\mathrm{cos}\alpha `$, instead of $`T=Mg`$ for a single wire. The elastic distance is however determined by the tension in each wire, $`\mathrm{\Delta }=\sqrt{EI/T}=\sqrt{2EI/Mg\mathrm{cos}\alpha }`$. To keep the stress in the wires constant, the cross section area of a single supporting wire is twice the area of two each of two supporting wires. Thus, the effective quality factor determining the thermal noise for the displacement thermal noise $`x^2(f)`$ has a smaller dilution factor of $`\mathrm{\Delta }/L=1/326`$, instead of 1/231 for a single wire. This is the well-know effect of reducing thermal noise by increasing the number the wires. The dilution factor for pitch is $`1/14`$, considerably higher than the dilution factor for yaw, $`1/44`$.
The displacement thermal noise at 160 Hz is $`\sqrt{x^2(160\mathrm{H}\mathrm{z})}=1.1\times 10^{20}\mathrm{m}/\sqrt{\mathrm{Hz}}`$, limiting the detector sensitivity to $`h=5.6\times 10^{24}/\sqrt{\mathrm{Hz}}`$. This is expected to be lower than the thermal noise due to the internal modes of the mirror mass, not considered here . The pendulum thermal noise could be reduced by a factor $`\sqrt{2}`$, or about 40%, if the beam spot was positioned at the optimal position on the mirror. Since pendulum thermal noise is not the dominant source noise, but the detectorsโ shot noise would increase due to diffraction losses, it is not advisable for LIGO to proceed this way. However, these considerations should be taken into account for future detectors, where thermal noise may be a severe limitation at low frequencies.
The pitch angular noise at 160 Hz, is $`\sqrt{\theta ^2(f)}=8.9\times 10^{19}\mathrm{rad}\sqrt{\mathrm{Hz}}`$. Its contribution to the sensed motion has to take into account the coupling with displacement, and we will do this in detail inthe last section.
### B Vertical, transverse displacement and roll
We are now concerned with the mirror motion in its $`y,z`$ and $`\psi `$ degrees of freedom. The potential energy is $`PE_i=(1/2)_0^L\left(T(w_{i}^{}{}_{}{}^{})^2+EI(w_{i}^{\prime \prime }{}_{}{}^{})^2+EA(w_{i}^{}{}_{}{}^{})^2\right)๐s`$ for each wire, plus $`T(h\mathrm{cos}\alpha b\mathrm{sin}\alpha )\psi ^2`$. Notice that due to the tilting of the wires, the โtransverseโ $`w_{}`$ and โaxialโ $`w_{}`$ directions are not $`y`$ and $`z`$, but rotations of these directions by the wire tilt angle $`\alpha `$. We define $`w_{i}^{}{}_{}{}^{}`$ for each wire pointing โoutโ (and thus in opposite directions if $`\alpha =0`$), and $`w_{i}^{}{}_{}{}^{}`$ pointing down along the wire, from top to bottom.
The boundary conditions at top are $`w_{i}^{}{}_{}{}^{}(0)=0,w_{i}^{}{}_{}{}^{}(0)=0,w_{i}^{}{}_{}{}^{}(0)=0`$, and at the bottom, $`w_{i}^{}{}_{}{}^{}(L)=\pm (y\mathrm{cos}\alpha d\psi )+z\mathrm{sin}\alpha `$, $`w_{i}^{}{}_{}{}^{}(L)=\pm \psi `$, and $`w_{i}^{}{}_{}{}^{}(L)=\pm (y\mathrm{sin}\alpha c\psi )z\mathrm{cos}\alpha `$, where we defined two new distances $`c=h\mathrm{sin}\alpha +b\mathrm{cos}\alpha `$ and $`d=h\mathrm{cos}\alpha b\mathrm{sin}\alpha `$. If we define as earlier, sums and differences of the two wires shape functions, $`w_\pm (s)=(w_1(s)\pm w_2(s))/2`$, then the equations of motion for the mirror degrees of freedom, when subject to external forces $`F_y,F_z`$ and a torque $`N_x`$, are
$`F_y`$ $`=`$ $`M\omega ^2y+2(Tw_{}^{}{}_{}{}^{}(L)EIw_{}^{\prime \prime \prime }{}_{}{}^{}(L))\mathrm{cos}\alpha `$
$`+2EAw_{}^{}{}_{}{}^{}\mathrm{sin}\alpha `$
$`F_z`$ $`=`$ $`M\omega ^2z+2(Tw_{+}^{}{}_{}{}^{}(L)EIw_{+}^{\prime \prime \prime }{}_{}{}^{}(L))\mathrm{sin}\alpha `$
$`2EAw_{+}^{}{}_{}{}^{}\mathrm{cos}\alpha `$
$`N_x`$ $`=`$ $`J_x\omega ^2\psi 2EI(dw_{}^{\prime \prime \prime }{}_{}{}^{}(L)+w_{}^{\prime \prime }{}_{}{}^{}(L))`$
$`+2EAcw_{}^{}{}_{}{}^{}(L)`$
The solution for the wiresโ transverse motion $`w_{\pm }^{}{}_{}{}^{}(s)`$ satisfying the boundary conditions up to order $`e^{L/\mathrm{\Delta }}`$ and $`k\mathrm{\Delta }`$ are of the same form as in Eqn. 3, with top and bottom weights equal to
$`A_{}`$ $`=`$ $`(y\mathrm{cos}\alpha (d+\mathrm{\Delta })\psi )/D`$ (25)
$`B_{}`$ $`=`$ $`((y\mathrm{cos}\alpha d\psi )(cos(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))`$ (27)
$`\psi (\mathrm{sin}(kL)k\mathrm{\Delta }\mathrm{cos}(kL))/k)/D`$
$`A_+`$ $`=`$ $`z\mathrm{sin}\alpha /D`$ (28)
$`B_+`$ $`=`$ $`z\mathrm{sin}\alpha (cos(kL)+k\mathrm{\Delta }\mathrm{sin}(kL))/D`$ (29)
The axial wire motion is
$$w_{\pm }^{}{}_{}{}^{}=w_{\pm }^{}{}_{}{}^{}(L)\frac{\mathrm{sin}(k_zs)}{\mathrm{sin}(k_zL)}$$
(30)
where $`w_{}^{}{}_{}{}^{}(L)=y\mathrm{sin}\alpha c\psi `$ and $`w_{+}^{}{}_{}{}^{}(L)=z\mathrm{cos}\alpha `$. The wavenumber functions are $`k^2=\rho \omega ^2/T`$, $`k_z^2=\rho \omega ^2/EA`$.
Even though the tilting of the wires produces more complicated formulas than in the pendulum-pitch-yaw case, the equations for the vertical motion decouple from the equations from the transverse pendulum displacement and roll, similar to yaw decoupling from pendulum and pitch. As before, if the wires move in phase (transverse or axially or both), they produce only vertical motion; but if they move in opposition, they produce side to side motion plus rotation around the optical axis. We analyze the two decoupled systems separately.
Vertical thermal noise. Once we have solved the wire shape (from Eqns3,29, and 30), we can write the equation of motion of the wire vertical displacement as
$`F_z=M\omega ^2z`$ $`+2({\displaystyle \frac{EA}{L}}{\displaystyle \frac{k_zL}{\mathrm{tan}(k_zL)}}\mathrm{cos}^2\alpha `$
$`+Tk{\displaystyle \frac{\mathrm{cos}(kL)+k\mathrm{\Delta }\mathrm{sin}(kL)}{\mathrm{sin}(kL)2k\mathrm{\Delta }\mathrm{cos}(kL)}}\mathrm{sin}^2\alpha )z`$
or
$$F_z=M\omega ^2z+(K_T\mathrm{sin}^2\alpha +K_E\mathrm{cos}^2\alpha )z$$
where we defined $`K_T`$ as the spring constant that was used in the pendulum-pitch case, and $`K_E2EA/L`$ the spring constant of the wire. For LIGO parameters, and for usual wires, $`K_T/K_E1`$.
If the wires are not tilted and $`\alpha =0`$, we recover the simple case of vertical modes of a mirror hanging on a single wire. The restoring force is elastic and proportional to $`E`$, so there is no dilution factor. The term added because of the wire tilting is a gravitational restoring force, much smaller than the elastic restoring force. Since it is also mostly real when considering a complex Young modulus $`E`$, it will not change significatively the loss terms, and thus the thermal noise. The wire tilting does add, however, the violin modes to the vertical motion, and it slightly decreases (by a factor $`\mathrm{cos}\alpha =0.97`$) the frequency of the lowest vertical mode.
The vertical thermal noise at 160 Hz is $`\sqrt{z^2(f)}=3.1\times 10^{18}\mathrm{m}/\sqrt{\mathrm{Hz}}`$, 260 times the pendulum thermal noise. This is due to the lower quality factor, and the higher mode frequency. However, vertical noise is sensed in the gravitational wave interferometer through the angle of the laser beam and the normal to the mirror surface, which is not less than the Earthโs curvature over 4km, (0.6 mrad). At the minimum coupling (0.3 mrad for each mirror in the 4km cavity), the contribution due to vertical thermal noise is 10% of the pendulum thermal noise. In advanced detectors, vertical modes are going to be at lower frequencies due to soft vertical supports, like in the suspensions used in the GEO600 interferometer, but the ratio of quality factors is just the mechanical dilution factor, so the contribution of vertical thermal noise to sensed motion will be of order $`3\times 10^4\times \sqrt{L/\mathrm{\Delta }}\times f_z/f_x3\times 10^4\times (EL^2/Mg)^{1/4}`$, not necessarily a small number!
Side pendulum and roll. The side motion and roll of the pendulum are not expected to appear in the interferometer signal, but it is usually the case that at least the high quality-factor resonances do appear through imperfect optic alignment. The equations for the system $`y,\psi `$ can be written as
$`(K_{yy}M\omega ^2)yK_{y\psi }\psi `$ $`=`$ $`F_y`$
$`(K_{\psi \psi }J_x\omega ^2)\psi K_{\psi y}y`$ $`=`$ $`N_x`$
with
$`K_{yy}`$ $`=`$ $`K_T\mathrm{cos}^2\alpha +K_E\mathrm{sin}^2\alpha `$ (31)
$`K_{\psi \psi }`$ $`=`$ $`K_Ec^2+K_Td(d+\mathrm{\Delta })`$ (33)
$`+2T(d+\mathrm{\Delta }){\displaystyle \frac{\mathrm{sin}(kL)k\mathrm{\Delta }\mathrm{cos}(kL)}{\mathrm{sin}(kL)2k\mathrm{\Delta }\mathrm{cos}(kL)}}`$
$`K_{y\psi }=K_{\psi y}`$ $`=`$ $`K_T(d+\mathrm{\Delta })\mathrm{cos}\alpha +K_Ec\mathrm{sin}\alpha `$ (34)
Within the (very good) approximation $`K_T/K_E1`$, we can prove that the eigenfrequencies are $`w_y^2=K_T/M,f_y=0.75`$ Hz and $`\omega _\psi ^2=K_Eb^2/J,f_\psi =19`$ Hz. The corresponding loss factors are $`\varphi _y\mathrm{\Delta }\varphi /2L`$ and $`\varphi _\psi \varphi `$.
If the wires are perfectly straight, the thermal noise of the side-to-side pendulum motion below the violin modes $`y_0^2(f)`$ is well approximated by that of a simple oscillator with eigenfrequency $`\omega _y`$ and a dilution factor $`\mathrm{\Delta }/2L`$. The thermal noise of the roll angular motion $`\psi ^2(f)`$ does not depend much on the wire tilt, and is well approximated (below violin modes) by the thermal noise of a simple oscillator with eigenfrequency $`w_r^2=2EA/J_xL`$ and quality factor equal to the free wireโs quality factor (there is no dilution factor).
For any small wire tilt, however, the spring constant $`K_{yy}`$ in Eqn.34 has a large contribution of $`K_E`$, and thus the fluctuations increase as $`y^2(f)y_0^2(f)\mathrm{cos}^2\alpha +(J_x/Mb)^2\psi ^2(f)\mathrm{sin}^2\alpha `$, as seen in Fig. 9. Since the roll eigenfrequency is higher than the side pendulum eigenfrequency, and its quality factor is lower, this is a significant increase in the thermal noise spectral density $`\psi ^2(f)`$, about a factor of 100 for LIGO parameters. However, the gravitationalw ave detectors are mostly immune to the roll degree of freedom, as we will see later.
Also, the tilting of the wires introduces violin modes harmonics of $`f_n^{(2)}=\sqrt{(}EA/\rho )/(2L)n\times `$6kHz apart from the usual harmonics $`f_n^{(1)}=\sqrt{(}T/\rho )/(2L)n\times 300`$Hz.
The violin modes that are most visible in the roll thermal noise are the harmonics of $`\sqrt{(}EA/\rho )/(2L)`$ (and are strictly the only ones present if the wires are not tilted).
### C Violin Modes
We have explored the relationship between quality factors of pendulum modes and the โeffectiveโ quality factor needed to predict the thermal noise of any given degree of freedom (seen only in sensitive interferometers). For the most important longitudinal motion, weโve seen that the effective quality factor is approximately equal to $`Q_wL/\mathrm{\Delta }`$, where $`Q_w=1/\varphi _w`$ is the quality factor of the free wire, related to the imaginary part of the Young modulus. Under ideal conditions where the pendulum mode is far away from the pitch mode in frequency, its quality factor is close to the effective quality factor, but as we have seen, the errors may be as large as 50%.
The violin modes, approximately equal to $`f_n=n\sqrt{T/\rho }/2L`$, appear in the horizontal motion of the pendulum in both directions, along the optical axis and transverse to it. The violin modes show some anharmonicity, as pointed in ,with the frequencies slightly higher than $`n\sqrt{T/\rho }/2L`$, and the quality factors degrading with mode number. The complex eigenfrequencies are the solutions to the equation
$$kL=n\pi +\mathrm{arctan}(2k\mathrm{\Delta }/(1(k\mathrm{\Delta })^2))$$
with $`k^2=\rho \omega ^2/T`$. This means that if we measure the quality factors of violin modes, and they follow the predicted anharmonic behavior in both directions, we can assume the losses are only limited by wire losses. We can then predict the thermal noise in the gravitational wave band with more confidence, having also consistency checks with the quality factors measured at the pendulum modes. The thermal amplitude of the peaks at the violin modes follows a simple $`1/f^{5/2}`$ law, corrected by the change in longitudinal Qs. We show all these features in Fig.12.
### D Total Pendulum Thermal Noise in LIGO I
The right way to calculate the total thermal noise observed in the interferometer signal is to calculate the pendulum response to an applied oscillating force in the direction of the laser beam. The pendulum responds to a force in all its six degrees of freedom, but the motion we are sensitive to is the motion projected on the laser beamโs direction.
If the laser beam is horizontal, and its direction passes through the mirror center of mass, it will be sensitive to only longitudinal displacement and pitch motion. If the beam is not horizontal, and for example is tilted up or down by an angle $`\gamma `$, (but still going through the center of mass), we imagine an applied force $`F`$ applied in the beamโs direction, with a horizontal component $`F\mathrm{cos}\gamma `$ and a vertical component $`F\mathrm{sin}\gamma `$. The motion we are interested in is $`x\mathrm{cos}\gamma +z\mathrm{sin}\gamma `$, since it is the direction sensed by the laser beam. The admittance we need to calculate is then $`Y=i\omega (x\mathrm{cos}\gamma +z\mathrm{sin}\gamma )/F=i\omega ((x/F_x)\mathrm{cos}\gamma +(z/F_z)\mathrm{sin}\gamma )=i\omega (Y_x\mathrm{cos}^2\gamma +Y_z\mathrm{sin}^2\gamma )`$, where $`Y_x`$ is the response of the horizontal displacement to a horizontal force and $`Y_z`$ is the vertical response to a vertical force. A force applied in the beamโs direction, with magnitude $`F_0`$, will have components in all 3 axes, and torques around all 3 axes too: $`\stackrel{}{F}=F_0(\alpha _x,\alpha _y,\alpha _z)`$ and $`\stackrel{}{N}=F_0(D_\psi ,D_\theta ,D_\psi )`$. The motion we are interested is, in general, $`\chi =\alpha _xx+\alpha _yy+\alpha _zz+D_\theta \theta +D_\varphi \varphi +D_\psi \psi `$. If the motion in all 6-DOF was uncoupled, then each dof responds to just one component of the force or the torque, and the admittance we need would be just the sum of the admittances, each weighted by the square of a factor $`\alpha _i`$ or a distance $`D_i`$. However, only $`z`$ and $`\varphi `$ are decoupled dof from the rest, and $`x,\theta `$ and $`y,\psi `$ form two coupled systems, for which we need to solve the response to a forces and torques together.
We define an admittance $`Y_{x\theta }=i\omega (x/N_y)`$ as the admittance of displacement $`x`$ to an applied torque $`N_x`$, and $`Y_{\theta _x}=i\omega (\theta /F_x)`$ as the admittance of pitch to an applied horizontal force , and similar quantities $`Y_{y\psi },Y_{\psi ,y}`$. Then, the total thermal noise sensed by the laser beam is
$`\chi ^2(f)`$ $`={\displaystyle \frac{4k_BT_0}{\omega ^2}}\mathrm{}`$ $`(\alpha _x^2Y_{xx}+\alpha _xD_\theta (Y_{x\theta }+Y_{\theta x})+D_\theta ^2Y_{\theta \theta }`$
$`+\alpha _y^2Y_{yy}+\alpha _yD_\psi (Y_{y\psi }+Y_{\psi y})+D_\psi ^2Y_{\psi \psi }`$
$`+\alpha _z^2Y_{zz}+D_\varphi ^2Y_{\varphi \varphi })`$
or
$`\chi ^2(f)`$ $`=`$ $`\alpha _x^2x^2(f)+D_\theta ^2\theta ^2(f)+\alpha _xD_\theta {\displaystyle \frac{4k_BT_0}{\omega ^2}}\mathrm{}(Y_{x\theta }+Y_{\theta x})`$
$`+\alpha _y^2y^2(f)+D_\psi ^2\psi ^2(f)`$
$`+\alpha _yD_\psi {\displaystyle \frac{4k_BT_0}{\omega ^2}}\mathrm{}(Y_{y\psi }+Y_{\psi y})`$
$`+\alpha _z^2z^2(f)+D_\varphi ^2\varphi ^2(f)`$
so it is a weighted sum in quadrature of the thermal noise of different degrees of freedom, plus some cross-terms. These terms may be negative, so it is possible to choose an optimal set of parameters to minimize the sensed motion, as shown in . The weighting factors and distances are
$`\alpha _x`$ $`=`$ $`\mathrm{cos}(\gamma _y)\mathrm{cos}(\gamma _z)`$
$`\alpha _y`$ $`=`$ $`\mathrm{sin}(\gamma _y)\mathrm{cos}(\gamma _z)`$
$`\alpha _z`$ $`=`$ $`\mathrm{sin}(\gamma _z)`$
$`D_\theta `$ $`=`$ $`R\mathrm{sin}(\gamma _z)+dz\mathrm{cos}(\gamma _y)\mathrm{cos}(\gamma _z)`$
$`D_\varphi `$ $`=`$ $`\mathrm{cos}(\gamma _z)(d_y\mathrm{cos}(\gamma _y)R\mathrm{sin}(\gamma _y))`$
$`D_\psi `$ $`=`$ $`d_y\mathrm{sin}(\gamma _z)+d_z\mathrm{sin}(\gamma _y)\mathrm{cos}(\gamma _z)`$
Typical distances $`d_y,d_z`$ are few mm at most, and typical angles $`\gamma _y`$are in the order of microradians, since an angle of such magnitude produces displacements of the order of millimeters at the beam at the other end of the arm, 4km away. However, the angle $`\gamma _z`$ cannot be less than half the arm length divided by the curvature of Earth, or $`\gamma _z3\times 10^4`$. We show in Fig13 the thermal noise sensed by a beam with $`d_y=d_z=5`$mm , $`\gamma _z=3\times 10^4`$ and $`\gamma _y=1\mu `$rad.
## IV Results and conclusions
We have shown several general results, and then calculated predicted thermal motions for LIGO suspensions. First, we showed that if there is a single oscillator with two sources of potential energy, one with a spring constant $`K_g`$ and a dominant real part, and the other with a complex spring constant $`K_e`$, with a dominant loss factor, then the dilution factor $`K_e/K_g`$ gives us the ratio between the oscillatorโs quality factor (determining its thermal noise spectral density) and the loss factor of $`K_e`$. However, the elastic loss $`\varphi _e=\mathrm{}K_e/\mathrm{}K_e`$ might itself have a small dilution factor with respect to the loss factor of the Young modulus, for example if $`K_e\sqrt{E}`$, then there is a dilution factor of 1/2.
We also show that when two or more degrees of freedom are coupled, the measurable quality factors at resonance may not be as useful to predict the โeffectiveโ quality factor used in the thermal noise spectral density, unless the eigenfrequencies are far from each other.
We showed that by using approximations of order $`e^{L/\mathrm{\Delta }}`$, we can easily obtain wire shapes and equations of motion for the 6 degrees of freedom of the mirror, as a function of applied oscillating forces and torques. We think that this method will be most useful when applied to multiple pendulum systems such as those used in GEO600 and planned for advanced LIGO detectors . However, even in the simple pendulum case this approach allows us to calculate the gravitational and elastic potential energy as a linear energy density along the wire, and the total energy. We showed that, for an applied horizontal force, the gravitational potential energy ($`(1/2)Tw^2๐s`$) is homogeneously distributed along the wires, while the elastic potential energy ($`(1/2)EI(w^{\prime \prime })^2๐s`$) is concentrated at the top and bottom, but in different proportions depending on the frequency of the applied force (Fig. 5. We also calculate the ratio of total elastic potential energy to gravitational energy using the solutions for the wire shape, and show that this function of frequency corresponds to the dilution factor for the eigenmode loss factors as well as for the effective quality factor that allows us to calculate the thermal noise at gravitational wave frequencies (see Figs.1,4,8).
Applying our calculation of wire shapes and equations of motion that include elasticity to LIGO suspensions, we show in Figs 10 and 11 the resulting spectral densities of displacement and angular degrees of freedom of the mirror. More importantly, we show in Fig13 the resulting contribution of pendulum thermal noise to the LIGO sensitivity curve, assuming small misalignments in the sensing laser beam (5mm away from center of mass, 1$`\mu `$rad away from horizontal). As expected, the displacement degree of freedom is the one that dominates the contribution, but pitch noise contributes significantly (29% at 100 Hz) if the beam is 5mm above center. But this is not added in quadrature to the horizontal noise, which makes 81%: the coupled displacement-pitch thermal motion makes up 99% of the total thermal noise. A misplacement below the center of mass will reduce the observed thermal noise, as first noted in . The contribution of yaw thermal noise (11% at 100Hz) is smaller than that of pitch, but very comparable. The contribution of vertical noise due to the 4km length of the interferometer is 8% at 100 Hz. The side and roll motions are coupled, but the roll contribution dominates (due to the large angle $`\gamma _z`$ and the assumed $`d_z=5`$mm) and is 0.7%, much smaller than the contributions of pitch, yaw and vertical degrees of freedom. When added in quadrature, the total thermal noise is 23% higher than the contribution of just the horizontal thermal noise. However, if the vertical misplacement of the beam is 5mm below the center of mass, instead of above, the total contribution of thermal noise is 89% of the horizontal thermal noise of the center mass.
## V Acknowledgments
Much of this work was motivated by many discussions held with Jim Hough, Sheila Rowan and Peter Saulson, and I am very glad to thank their insights. I want to especially thank P. Saulson for carefully reading the original manuscript and making important suggestions. I also want to thank P. Fritschel and Mike Zucker, who first asked about thermal noise of angular modes in pendulums. This work was supported by NSF grants 9870032 and 9973783, and by The Pennsylvania State University.
|
warning/0006/hep-ph0006169.html
|
ar5iv
|
text
|
# Chiral symmetry breaking in the Wegner-Houghton RG approach
## 1 INTRODUCTION
The full understanding of the phase structure of QCD as the theory of strong interactions has become an important issue since the discovery of asymptotic freedom. The possibility that high temperature QCD could show different properties from the theory at zero temperature has been addressed already in . New features are also predicted for the theory at very high and intermediate baryon density , .
An essential role is played by the Chiral Symmetry which is supposed to be broken by the vacuum structure of the theory. Thus, before considering the high temperature phase transition with the Chiral Symmetry restoration, one has to deal with the problem of determining a nonzero order parameter which indicates the Chiral Symmetry Breaking (CSB) at zero temperature and density. A strong simplification is obtained by considering the effective theory described by a linear $`\sigma `$ model of scalar mesons coupled to an isospin multiplet of fermions with the same chiral symmetry group of the original action. Then CSB corresponds to a mexican hat-shaped effective potential and the order parameter is identified with the vacuum expectation value (VEV) of one of the scalar fields.
The infrared (IR) properties of a field theory such as the vacuum structure, can be analysed by means of the wilsonian renormalization group which generates a sequence of effective actions defined at some momentum scale $`k`$, starting from the original action, by integrating out in the latter all the modes with frequency higher than $`k`$. This leads to the construction of a differential flow equation for the $`k`$-dependent action, generally known as Exact Renormalization Group Equation (ERGE). In the past years many analytical and numerical approaches to the ERGE, applied to various theories, have been developed. Two recent extensive reviews of this topic are and a detailed ERGE analysis of the Chiral Phase Transition can be found in .
Here we shall consider just the simple case of the CSB at zero temperature employing the sharp cut-off version of the ERGE, namely the Wegner-Hougton equation . The cut-off dependence of the ERGE for the O(4) linear $`\sigma `$-model has been addressed in where, however only the flow of the scalar potential is considered while the Yukawa coupling is kept fixed. In addition to the Yukawa coupling flow, we are particularly interested in the behavior of the field renormalization. In fact, if our model is really an effective theory which, at some scale, can naturally replace the original QCD action, then at this scale one expects the formation of the mesonic bound states characterized by a very small wave function renormalization function. The reduction of QCD to the quark-meson action has been analysed before and a review of it is given in . As expected the value of the scalar field renormalization is vanishingly small at the mesonic bound state formation scale which is found between 0.60 and 0.63 GeV. Then, the flow of the field renormalization deserves particular attention.
## 2 ANALYSIS AND RESULTS
The explicit form of the euclidean action at a scale $`k`$ is ($`\rho \sigma ^2+\stackrel{}{\pi }^2`$)
$$S_k=d^4x\left(\frac{z_{l}^{}{}_{k}{}^{}}{2}_\mu \sigma _\mu \sigma +\frac{z_{t}^{}{}_{k}{}^{}}{2}_\mu \stackrel{}{\pi }_\mu \stackrel{}{\pi }+U_k(\rho )+\underset{c}{\overset{N_c}{}}\overline{\psi }_c\left(i\gamma +g_k\sigma +ig_k\gamma ^5\stackrel{}{\pi }\stackrel{}{\tau }\right)\psi _c\right)$$
(1)
$`\psi _c`$ is a flavor doublet and the summation over the index $`c`$ is extended to $`N_c`$ colors. No CSB fermion mass term is included and the action is fully invariant under chiral transformations. The presence of CSB depends on the shape of the O(4) symmetric potential $`U_k`$ in the limit $`k0`$. $`z_{l}^{}{}_{k}{}^{}`$ and $`z_{t}^{}{}_{k}{}^{}`$ are the $`\sigma `$ and $`\pi `$ renormalization functions. As long as $`U_k`$ has zero vacuum expectation value (VEV), we choose the following parametrization $`U_k=(1/2)m_k^2\rho +(\lambda _k/24)\rho ^2`$ and consider the flow of the two parameters $`m_k^2`$ and $`\lambda _k`$ neglecting all the irrelevant operators that could be generated during the flow. In the presence of a nonvanishing VEV $`\overline{\rho }_k`$, according to , we reparametrize the potential in terms of $`\overline{\rho }_k`$ and $`\lambda _k`$, recalling that $`m_k^2`$ can be expressed in term of these two parameters through the minimum condition. The purpose of this double choice is practical, we take the couple of parameters which corresponds to the simplest set of flow equations.
The Wegner-Houghton equation can be reduced to a set of coupled ordinary differential equations for the $`k`$ dependent parameters in (1), i.e. $`g_k`$, $`z_{l}^{}{}_{k}{}^{}`$, $`z_{t}^{}{}_{k}{}^{}`$, $`\lambda _k`$ and $`m_k^2`$ or $`\overline{\rho }_k`$. In particular the equations for the parameters in the potential are deduced in , and the one for $`g_k`$ can be obtained by an analogous procedure. The equations for $`z_{l}^{}{}_{k}{}^{}`$, $`z_{t}^{}{}_{k}{}^{}`$, (which coincide as long as $`\overline{\rho }_k=0`$, according to the chiral simmetry) are deduced following the procedure formulated in and already implemented in . In the following, the derivatives of $`z_{l}^{}{}_{k}{}^{}`$ and $`z_{t}^{}{}_{k}{}^{}`$ w.r.t. the fields, which are supposed to be small, are neglected.
The pion decay constant and the constituent fermion mass are two IR constraints for our equations. In fact, in the limit $`k0`$ we have $`\sqrt{z_{l}^{}{}_{k}{}^{}\overline{\rho }_k}ff_\pi 0.09`$ Gev and $`g_k\sqrt{\overline{\rho }_k}MM_q0.3`$ GeV, as initial conditions for the equations. If we start, according to , the flow at the UV scale $`k=0.6`$ GeV with a convex potential ($`m_k^2>0`$), then at a certain scale $`K_\chi `$, $`m_k^2`$ becomes negative generating a nonzero $`\overline{\rho }_k`$. In the following the three remaining initial conditions imposed are $`\overline{\rho }_k=0`$ at the CSB scale $`K_\chi `$ and a particular value for $`z_{l}^{}{}_{k}{}^{}`$ at the UV scale (the symmetry requires $`z_{l}^{}{}_{k}{}^{}=z_{t}^{}{}_{k}{}^{}`$ at the UV scale).
Figure LABEL:fig:uno shows, above $`K_\chi `$ (fixed at the value $`K_\chi =0.43`$ GeV), the flow of the scalar mass $`m`$ (the subscript $`k`$ is omitted) for two different initial values of $`z_l(=z_t)`$ and, below $`K_\chi `$, the corresponding $`f`$ and $`M`$. The curves for $`f`$ and $`M`$ in the two cases are almost superposed indicating insensitivity to the UV condition on $`z_l`$. The value of $`K_\chi `$ is suggested by an IR stability criterion. In order to use a particular value of quark mass, $`M_q`$, as the initial condition for the running quantity $`M`$ defined above, $`M`$ must be practically flat around $`k=M_q`$. The same criterion should apply to $`f_\pi `$ but, since $`f_\pi `$ is smaller than $`M_q`$, we find that $`f`$ at the scale $`k=f_\pi `$ is already a stable parameter. After fixing the IR condition $`M(k=M_qฯต)=M_q`$ with $`ฯต=0.03`$ GeV, we have considered the ratio $`\mathrm{\Delta }=(M(k=M_q)M_q)/M_q`$, plotted vs. $`K_\chi `$ in Figure LABEL:fig:due for two different values of $`M_q`$. The stability requirement indicates that $`K_\chi `$ should not be smaller than $`0.4`$ Gev.
In Figures LABEL:fig:tre and LABEL:fig:quattro the other running parameters are displayed for the same two initial values of $`z_l`$ that have been used in Figure LABEL:fig:uno. It should be noted the small running of $`g`$ and $`\lambda `$, in the range of $`k`$ considered, and their large values which are clearly nonperturbative. As shown in Figure LABEL:fig:quattro, when going from the UV to the IR scale, the field renormalizations grow and, below $`K_\chi `$, $`z_l`$ and $`z_t`$ increase differently. Since reasonably we expect, for the pion physics at scales around $`k=0.3`$ GeV, field renormalization functions close to the unity, we must require very small values at $`k=0.6`$ GeV. Namely we used $`z_l=z_t=0.1`$ and $`z_l=z_t=0.01`$ in the two cases here considered in Figures LABEL:fig:uno,LABEL:fig:tre, LABEL:fig:quattro. Higher curves for $`m,\lambda ,g`$ correspond to the former case, lower curves to the latter.
The approximation here employed for the ERGE has been sufficient for exploring a nonperturbative region (note that, as explained in , the renormalized couplings which include the field renormalization effect, are larger than those in Figure LABEL:fig:tre) and it has yielded a very small (vanishing) scalar field renormalization around $`0.6`$ GeV. We hope to extend our analysis to the finite temperature case.
|
warning/0006/cond-mat0006300.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The phenomenon of the optical activity (OA) has two aspects which arise from the interaction of the radiation with matter - dispersive and absorptive. These aspects are the optical rotatory dispersion (ORD) and the circular dichroism (CD). It is well known that ORD is the dependency of the rotation of the linear polarized light per unit length on the frequency $`\omega `$ or on the wave length $`\lambda `$. The CD is the ellipticity per unit length of the wave getting off the crystals. In the difference from the ORD, which is nonzeroth in a wide frequency region, the CD is nonzeroth only in a very narrow frequency region in the absorption region.
The important group of optically active crystals are the crystals with screw axis of symmetry belonging to the space groups of symmetry $`D_3^4`$ and $`D_3^6`$. The typical representatives of these crystals are $`\alpha `$-quartz, cinnabar, tellurium, selen, camphor, benzil. The optical activity of these crystals due to asymmetrical originating of the crystal structure because the molecules or atoms forming the crystals are not optically active. Camphor is the exception because his molecules are optically active so that his optical activity has two sides - molecular and crystalline.
In the past the OA of crystals was studied by more authors and their works are based on the different theories. We can introduce the theory of excitons , the theory of coupled oscillators or Lagrangian formalism . But it is known that the theory of coupled oscillators gives the results which are very good applicable in the fitting of the experimental data of OA. The model of coupled oscillators was for the first time used by Chandrasekhar which has applicated the Kuhn model of two coupled oscillators which represent the smallest unit of the optical active crystal. He has used the following model: The first oscillator lies in the plane $`z=0`$ and its position is given by direction cosines $`\alpha ,\beta ,\gamma `$. The second oscillator lies in the plane $`z=d`$ and its direction of vibration is turned by the angle $`\theta `$ around the $`z`$ axis which is parallel to the crystal axis $`c`$. Both oscillators lie on the helix which is given by the crystal structure. The $`z`$ axis (crystal axis $`o`$) has the direction of the propagation of the linear polarized electromagnetic wave. Both coupled oscillators forming one compound oscillator are identical. We denote the coupling constant between oscillators as $`Q_1`$. Chandrasekhar further has assumed that the number of compound oscillators in the volume unit is $`N/2`$, where $`N`$ is the number of single oscillators in the volume unit. The influence of the interactions between oscillators belonging to the other helices is neglected because of another type of couplings.
Becouse of the interaction between single oscillators in the compound oscillator the natural frequency $`\omega _0`$ is split into two frequencies $`\omega _1`$ and $`\omega _2`$ of normal modes of vibrations. For these normal modes Chandrasekhar has solved the dispersion theory of refractive indices for the propagation of the left and the right circularly polarized wave into which the linear polarized wave is split in the optically active medium. It is well known that the medium containing coupled oscillators is optically active and therefore it must be characterized by the different refractive indices $`n_l`$ and $`n_r`$ for the left and the right circularly polarized wave. Chandrasekhar has solved the dispersion theory semiclasically, he has introduced into results the oscillator strengths but he has assumed that the oscillator strengths of the normal modes of vibrations are the same.
Chandrasekhar has solved his model of optical activity for the $`\alpha `$-quartz. Because the CD is not measured for the $`\alpha `$-quartz and therefore nor the ORD in the absorption region Chandrasekhar has assumed the oscillators as undamped.
Later V. Vyลกรญn has removed Chandrasekharโs simplifications in the approach of the oscillator strengths of the normal modes of vibrations and he has obtained for the ORD which is denoted by $`\rho (\omega )`$ the formula
$$\rho (\omega )=\frac{\pi Nde^2\omega ^2}{mc^2}\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta \left[\frac{f_{g_1}}{\omega _1^2\omega ^2}+\frac{f_{q_2}}{\omega _2^2\omega ^2}\right],$$
(1)
where $`f_{q_1}`$ and $`f_{q_2}`$ are the oscillator strengths of the normal modes of vibrations, $`m`$ is the mass of the single oscillator, $`e`$ his electric charge and $`c`$ is the velocity of the light. The frequencies of the normal modes are $`\omega _1^2=\omega _0^2+Q_1`$, $`\omega _2^2=\omega _0^2Q_1`$. Because the splitting of the frequencies of the normal modes is very small we can (1) rewrite in the form containing the natural frequency of single oscillators
$$\rho (\omega )=\frac{\pi Nde^2\omega ^2}{mc^2}\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta \left[\frac{f_{q_2}f_{g_1}}{\omega _0^2\omega ^2}+\frac{Q_1\left(f_{q_1}+f_{q_2}\right)}{\left(\omega _0^2\omega ^2\right)^2}\right].$$
(2)
We see that the ORD based on the two coupled oscillator model leads in general to the two member formula. The first term on the right side of eq. (2) is known as Drudeโs term and the second term is known as Chandrasekharโs term. This is problem because we donโt know the oscillator strengths for the real crystals. We can solve the oscillator strengths only in approximations. For example in the approximation of the linear harmonic oscillator $`f_{q_1}=f_{q_2}=f_0`$ (the Chandrasekhar approximation) the formula (2) leads only to the Chandrasekhar formula
$$\rho (\omega )=\frac{2\pi Nde^2Q_1f_0}{mc^2}\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta \frac{\omega ^2}{\left(\omega _0^2\omega ^2\right)^2}.$$
(3)
On the other hand we can solve the oscillator strengths in the Heitler - London approximation which gives the result
$$\frac{f_{q_1}}{\omega _1}=\frac{f_{q_2}}{\omega _2}=\frac{f_0}{\omega _0}$$
(4)
and then ORD is given by the formula
$$\rho (\omega )=\frac{\pi Ne^2dQ_1f_0}{mc^2}\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta \frac{\omega ^2\left(\omega _0^2+\omega ^2\right)}{\omega _0^2\left(\omega _0^2\omega ^2\right)^2}.$$
(5)
We note that the formula (5) is the Agranovich type which was derived by the excitonโs theory.
The formula (2) or its special forms (3) and (5) was used in many works to describe the experimental data of ORD of the crystals, for example of the tellur , of the $`\alpha `$-quartz or of the crystal $`Bi_{12}GeO_{20}`$ and without exception with good results. On the other hand we think that the Chandrasekhar model still contains some simplifications which we want to solve in this paper.
The Chandrasekhar model assumes that $`N/2`$ coupled oscillators, where $`N`$ is the number of single oscillators, exist in the volume unit. But it means that we have $`N/2`$ isolated compound oscillators in the volume unit. This idea desribes for example the system of randomly oriented isolated non-interacting molecules. But in the crystals all adjacent molecules or atoms interact on the helices. It means that the second oscillator in the any compound oscillator is at the same time the first oscillator in the next compound oscillator. It is the first question how the results of the Chandrasekhar model will change in the case when we include all couplings between all adjacent oscillators into this model.
The second question follows from the fact that the two coupled oscillators model neglect all other couplings between oscillators on one helix. It is possible in the case when all oscillators lie on the line because the value of the coupling constant decreases with third power of the distance. But it is not true in the real crystals where the diameter and the step of the helix are comparable. In these cases we must include also the couplings between even and odd oscillators on the helices. Further in the crystals with the space groups of symmetry $`D_3^4`$ and $`D_3^6`$ the vibration directions of the adjacent oscillators contain the angle $`\theta =120deg`$. The vibration directions of the even or odd oscillators contain the angle $`\theta =240deg`$. From the formulae (2), (3) or (5) we see that the sense of ORD depends on the value of $`\mathrm{sin}\theta `$. It means that the couplings between even and odd oscillators on the helices have the opposite effect on the ORD in compare with the couplings between adjacent oscillators. We note that in the crystals with space groups of symmetry $`D_3^4`$ and $`D_3^6`$ the couplings between the first and the fourth oscillators which are randomly oriented donโt have the influence on the OA and then the other couplings are really neglectable.
## 2 The influence of the couplings between all adjacent oscillators
We can discuss this problem by means of the results of our previous paper . In this paper we have solved the application of the ORD Chandrasekhar model of three coupled oscillators. We have added to the Chandrasekhar model of two coupled oscillators the third oscillator which lies on the helix in the plane $`z=d`$ and its vibration direction is turned by the angle $`\theta `$ around $`z`$ axis with respect to the oscillator in the plane $`z=0`$. We include also as Chandrasekhar only the couplings between adjacent oscillators. The number of these compound oscillators in the volume unit is $`N/3`$.
The solving method is also the same as Chandrasekharโs with the exception that the natural frequency $`\omega _0`$ of each oscillators is split into three frequencies $`\omega _1`$, $`\omega _2`$ and $`\omega _3`$ and also the ORD result contains three terms regards to (1) that is
$`\rho (\omega )`$ $`=`$ $`{\displaystyle \frac{2\pi Nde^2\omega ^2}{3mc^2}}\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta `$ (6)
$`\times \left[{\displaystyle \frac{\left(\sqrt{2}+2\mathrm{cos}\theta \right)f_{q_1}}{\omega _1^2\omega ^2}}+{\displaystyle \frac{4f_{q_2}\mathrm{cos}\theta }{\omega _2^2\omega ^2}}+{\displaystyle \frac{\left(\sqrt{2}2\mathrm{cos}\theta \right)f_{q_3}}{\omega _3^2\omega ^2}}\right],`$
where $`\omega _1^2=\omega _0^2+\sqrt{2}Q_1`$, $`\omega _2^2=\omega _0^2`$ and $`\omega _3^2=\omega _0^2\sqrt{2}Q_1`$. We can rewrite this result with the natural frequency of oscillators $`\omega _0`$ but this result is involved too. We express this result only in the approximations of the oscillators strengths. In the Chandrasekhar linear harmonic oscillator approximation $`f_{q_1}=f_{q_2}=f_{q_3}=f_0`$ we get
$$\rho (\omega )=\frac{8\pi Nde^2Q_1f_0\omega ^2}{3mc^2}\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta \left[\frac{1}{\left(\omega _0^2\omega ^2\right)^2}\frac{2Q_1\mathrm{cos}\theta }{\left(\omega _0^2\omega ^2\right)^3}\right].$$
(7)
We see that the second term in the square brackets is very small with respect to the first one. It contains the coupling constant $`Q_1`$ that is very small and the denominator of this term is in the frequency region far from absorption much greater than the denominator of the first term. From this reason the second term can be neglected.
In the Heitler - London approximation that is given by the extended relation (4) we have
$$\rho (\omega )=\frac{4\pi Nde^2Q_1f_0\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta }{3mc^2}\left[\frac{\omega ^2\left(\omega _0^2+\omega ^2\right)}{\omega _0^2\left(\omega _0^2\omega ^2\right)^2}\frac{2Q_1\mathrm{cos}\theta }{\left(\omega _0^2\omega ^2\right)^3}\right]$$
(8)
and we can neglect again the second term in the square brackets. But then the results of two and three coupled oscillators are very similar, they differ only in the multiplicative constant $`4/3`$. We see the sense of this constant if we compare the results for one compound oscillator. We must divide the results of two coupled oscillators model (3) and (5) by $`N/2`$ (the number of compound oscillators in the volume unit) and the results of the three coupled oscillators model (7) and (8) by $`N/3`$. By the comparing of corresponding results we see that one triad of the coupled oscillators has the twofold effect on the ORD than one couple. From this it follows that the effects of couplings between adjacent oscillators are aditive. If we want to include all couplings in the two coupled oscillators model we must take as the number of coupled oscillators $`N`$ (in the case $`N\mathrm{}`$) because the number of couplings is the same as the number of single oscillators. In the case of three coupled oscillators model we take N/2 as the number of compound oscillators in the volume unit etc. Then both models of two and three coupled oscillators give the same results in the case of the including of the couplings between adjacent oscillators only. The mistake of this conclusion is given by the neglecting of the second terms in the square brackets in (7) and (8) but we can assume that in the cases of small couplings in the crystals the mistake is neglectable.
From our conclusions in this section it follows that if we use our results in the approximation of experimental data of ORD of the crystals we get different values of unknown parameters in these formulae. For example the ORD of $`\alpha `$-quartz was approximated by Chandrasekharโs and Vyลกรญnโs results of the two coupled oscillators model. The unknown parameters were the oscillator strengths of the normal modes of vibrations. The values of the oscillator strengths were in the results that are assumed as true smaller as the number of valence shell electrons . In the other results the values of the oscillator strengths were greater. But after our correction of the results of the two coupled oscillators model we get for all oscillator strengths a half of values and it would be difficult to say what results of the approximations of the experimental data are true and what the values of all oscillator strengths will be too small. It is the next reason for the assertion that we must include also the couplings between even and odd oscillators which due to its opposite effect on the ORD return for example in the case of $`\alpha `$-quartz the results of the approximations of oscillator strengths to the values that they have in .
## 3 The influence of the couplings between even and odd oscillators
The smallest system in which we can include the coupling between even or odd oscillators in the helix is the system of three coupled oscillators if we include also the coupling between the first and the third oscillator. The scheme of this system is the same as in the previous section. We characterize the coupling between the first and the third oscillator by the coupling constant $`Q_2`$. Regarding to the previous section we use now the value $`N/2`$ as the number of compound oscillators where $`N`$ is the number of single oscillators in the volume unit. It means that the last single oscillator in any compound oscillator is the first single oscillator in the next compound oscillator on the helix.
The solving ORD of this system by the Chandrasekhar semiclassical model is described in . We get for the ORD the three member formula
$`\rho (\omega )`$ $`=`$ $`{\displaystyle \frac{4\pi Nde^2\omega ^2}{mc^2}}\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta `$ (9)
$`\times \left[{\displaystyle \frac{\frac{\left(2\mathrm{cos}\theta +A_1\right)}{2+A_1^2}f_{q_1}}{\omega _1^2\omega ^2}}+{\displaystyle \frac{f_{q_2}\mathrm{cos}\theta }{\omega _2^2\omega ^2}}+{\displaystyle \frac{\frac{\left(2\mathrm{cos}\theta +A_3\right)}{2+A_3^2}f_{q_3}}{\omega _3^2\omega ^2}}\right],`$
where $`\omega _1^2=\omega _0^2+\frac{Q_2+\sqrt{8Q_1^2+Q_2^2}}{2}`$, $`\omega _2^2=\omega _0^2Q_2`$ and $`\omega _3^2=\omega _0^2+\frac{Q_2\sqrt{8Q_1^2+Q_2^2}}{2}`$ and $`A_1`$, $`A_3`$ are $`A_1=\frac{Q_2+\sqrt{8Q_1^2+Q_2^2}}{2Q_1}`$, $`A_3=\frac{Q_2\sqrt{8Q_1^2+Q_2^2}}{2Q_1}`$. We see that by the expression of this result regarding to the natural frequency we obtain complicated result and its form is also described in . Let us again be interested in the form of the result after the approximations of the oscillator strengths. If we neglect as in previous section all the terms containing $`1/(\omega _0^2\omega ^2)^3`$ we get in the Chandrasekhar approximation the result
$$\rho (\omega )=\frac{4\pi Nde^2f_0\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta \left(Q_1+2Q_2\mathrm{cos}\theta \right)}{mc^2}\frac{\omega ^2}{\left(\omega _0^2\omega ^2\right)^2}$$
(10)
and in the Heitler - London approximation
$$\rho (\omega )=\frac{2\pi Nde^2f_0\left(\alpha ^2+\beta ^2\right)\mathrm{sin}\theta \left(Q_1+2Q_2\mathrm{cos}\theta \right)}{mc^2}\frac{\omega ^2\left(\omega _0^2+\omega ^2\right)}{\omega _0^2\left(\omega _0^2\omega ^2\right)^2}.$$
(11)
If we compare these results with the results (7) and (8) after their discussion, it means after neglecting the terms with $`1/(\omega _0^2\omega ^2)^3`$ and after their correction by including all couplings between adjacent oscillators (the coefficient $`N/3`$ is substituted by $`N/2`$), we see that the results (10) and (11) are different only in the coefficient $`Q_1+2Q_2\mathrm{cos}\theta `$ which substitutes the coupling constant $`Q_1`$ in the eqs. (7) and (8). In the crystals belonging to the space groups of symmetry $`D_3^4`$ and $`D_3^6`$ is $`\mathrm{cos}\theta =0.5`$ and the coefficient gives the value $`Q_1Q_2`$. The couplings between even and odd oscillators have indeed the opposite efect on the ORD. Besides we include in our three coupled oscillators model all couplings between adjacent oscillators on the helices but only a half of couplings between even and odd oscillators. Nevertheless the influence of couplings with couplings constants $`Q_1`$ and $`Q_2`$ has the same form. If we include all couplings between even and odd oscillators the influence of these couplings will be, of course, much greater but we donโt know the expression for this influence.
Even and odd coupled oscillators on the helices are not contacted and we cannot use the method of the contacted compound oscillators from the section 2. If we want to solve this problem exactly we should solve the model of $`\overline{N}`$ coupled oscillators where $`\overline{N}`$ is the number of all single oscillators on one helix and in this model we should include all couplings between adjacent, even and odd oscillators. This model is not, of course, really solvable.
From this reason we have proceeded by another way. We have solved further the models of four, five etc. coupled oscillators in which we have included the couplings between adjacent, even and odd oscillators and the results we have expressed only with the natural frequency $`\omega _0`$ and only in the approximations of the oscillator strengths of the normal modes of vibrations and only for the crystals with the space groups of symmetry $`D_3^4`$ and $`D_3^6`$. Based on these results we have found the tendency in the expression that contains the coupling coefficient $`Q_2`$ and we have looked the limit of this expression with increased number of oscillators in the models. We show now the solving of the four coupled oscillators model. The solving of other models is similar but more complicated and we discuss only the results.
### 3.1 The four coupled oscillator model of the optical rotatory dispersion of crystals
We extend the three coupled oscillators model by the fourth oscillator which lies on the helix in the plane $`z=2d`$ and his vibration direction is turned by the angle $`2\theta `$ around $`z`$ axis with respect to the vibration direction of the oscillator in the plane $`z=0`$. The coupling constant $`Q_1`$ describes the coupling between adjacent oscillators and the constant $`Q_2`$ describes the couplings between the first and the third and between the second and the fourth oscillators. The motion equations of the oscillators in the fields of the left and the right circularly polarized wave are
$$\begin{array}{c}\ddot{r}_1+\omega _0^2r_1+Q_1r_2+Q_2r_3=F_1^{l,r},\\ \ddot{r}_2+\omega _0^2r_2+Q_1r_1+Q_1r_3+Q_2r_4=F_2^{l,r},\\ \ddot{r}_3+\omega _0^2r_3+Q_2r_1+Q_1r_2+Q_1r_4=F_3^{l,r},\\ \ddot{r}_4+\omega _0^2r_4+Q_2r_2+Q_1r_3=F_4^{l,r},\end{array}$$
(12)
where $`r_1`$, $`r_2`$, $`r_3`$, $`r_4`$ are the displacements of oscillators from equilibrium, $`F`$ denotes the projection of functioning forces to the motion direction of oscillators divided by the masses $`m`$ of oscillators (electrons). In all terms in equation (12) and further in the sign $`\pm `$ hold the $`+`$ sign for the left and the $``$ sign for the right cicularly polarized wave. The electric field vector $`\stackrel{}{E}`$ has in our case the components
$$\begin{array}{c}E_x^{l,r}=E_0\mathrm{cos}(\omega tk_{l,r}z),\\ E_y^{l,r}=\pm E_0\mathrm{sin}(\omega tk_{l,r}z)\end{array}$$
(13)
and then we can express the forces $`F_\eta ^{l,r}(\eta =1,2,3,4)`$ as
$$\begin{array}{c}F_1^{l,r}=\frac{eE_0}{m}[(\alpha \mathrm{cos}\theta +\beta \mathrm{sin}\theta )\mathrm{cos}(\omega t+\varphi _{l,r})\\ \pm (\alpha \mathrm{sin}\theta +\beta \mathrm{cos}\theta )\mathrm{sin}(\omega t+\varphi _{l,r})],\\ F_2^{l,r}=\frac{eE_0}{m}(\alpha \mathrm{cos}\omega t\pm \beta \mathrm{sin}\omega t),\\ F_3^{l,r}=\frac{eE_0}{m}[(\alpha \mathrm{cos}\theta \beta \mathrm{sin}\theta )\mathrm{cos}(\omega t\varphi _{l,r})\\ \pm (\alpha \mathrm{sin}\theta +\beta \mathrm{cos}\theta )\mathrm{sin}(\omega t\varphi _{l,r})],\\ F_4^{l,r}=\frac{eE_0}{m}[(\alpha \mathrm{cos}2\theta \beta \mathrm{sin}2\theta )\mathrm{cos}(\omega t2\varphi _{l,r})\\ \pm (\alpha \mathrm{sin}2\theta +\beta \mathrm{cos}2\theta )\mathrm{sin}(\omega t2\varphi _{l,r})],\end{array}$$
(14)
where $`e`$ is the electron charge and $`\varphi _{l,r}=k_{l,r}=\frac{2\pi n_{l,r}d}{\lambda }=\frac{n_{l,r}\omega d}{c}`$ is a phase shift.
We may express relations (12) in the normal coordinates which are
$$\begin{array}{c}q_1=\frac{1}{\sqrt{2(1+A_1^2)}}(r_1+A_1r_2+A_1r_3+r_4),\\ q_2=\frac{1}{\sqrt{2(1+A_2^2)}}(r_1+A_2r_2+A_2r_3+r_4),\\ q_3=\frac{1}{\sqrt{2(1+A_3^2)}}(r_1+A_3r_2A_3r_3r_4),\\ q_4=\frac{1}{\sqrt{2(1+A_4^2)}}(r_1+A_4r_2A_4r_3r_4),\end{array}$$
(15)
where
$$\begin{array}{c}A_1=\frac{Q_1+\sqrt{5Q_1^2+8Q_1Q_2+4Q_2^2}}{2(Q_1+Q_2)},\\ A_2=\frac{Q_1\sqrt{5Q_1^2+8Q_1Q_2+4Q_2^2}}{2(Q_1+Q_2)},\\ A_3=\frac{Q_1\sqrt{5Q_1^28Q_1Q_2+4Q_2^2}}{2(Q_1+Q_2)},\\ A_4=\frac{Q_1+\sqrt{5Q_1^28Q_1Q_2+4Q_2^2}}{2(Q_1+Q_2)}.\end{array}$$
(16)
As a result of coupling the natural frequency $`\omega _0`$ of each oscillator splits into four characteristic frequencies of the normal modes of vibrations
$$\begin{array}{c}\omega _1^2=\omega _0^2+\frac{Q_1+\sqrt{5Q_1^2+8Q_1Q_2+4Q_2^2}}{2},\\ \omega _2^2=\omega _0^2+\frac{Q_1\sqrt{5Q_1^2+8Q_1Q_2+4Q_2^2}}{2},\\ \omega _3^2=\omega _0^2\frac{Q_1\sqrt{5Q_1^28Q_1Q_2+4Q_2^2}}{2},\\ \omega _4^2=\omega _0^2\frac{Q_1+\sqrt{5Q_1^28Q_1Q_2+4Q_2^2}}{2}.\end{array}$$
(17)
Now it is possible to rewrite the motion equations of each oscillator
$$\ddot{q}_\eta ^{l,r}+\omega _\eta ^2q_\eta ^{l,r}=R_{q_\eta }^{l,r},\eta =1,2,3,4,$$
(18)
where $`R_{q_\eta }^{l,r}`$ are the forces in the normal coordinates. For them we get in general expression
$$R_{q_\eta }^{l,r}=\frac{eE_0}{m}(a_{q_\eta }^{l,r})\mathrm{cos}(\omega t+\sigma _{q_\eta }^{l,r}),\eta =1,2,3,4$$
(19)
where $`\sigma _{q_\eta }^{l,r}`$ is a phase shift and for $`(a_{q_\eta }^{l,r})`$ we have derived
$$\begin{array}{c}(a_{q_1}^{l,r})^2=\frac{\alpha ^2+\beta ^2}{1+A_1^2}[1\pm 3\varphi _{l,r}\mathrm{sin}\theta 3\mathrm{cos}\theta +4\mathrm{cos}^3\theta 12\varphi _{l,r}\mathrm{cos}^2\theta sin\theta \\ 2A_1(1\mathrm{cos}\theta 2\mathrm{cos}^2\theta \pm \varphi _{l,r}\mathrm{sin}\theta \pm 4\varphi _{l,r}\mathrm{sin}\theta \mathrm{cos}\theta )\\ +A_1^2(1+\mathrm{cos}\theta \varphi _{l,r}\mathrm{sin}\theta )]\\ (a_{q_2}^{l,r})^2=\frac{\alpha ^2+\beta ^2}{1+A_2^2}[1\pm 3\varphi _{l,r}\mathrm{sin}\theta 3\mathrm{cos}\theta +4\mathrm{cos}^3\theta 12\varphi _{l,r}\mathrm{cos}^2\theta sin\theta \\ 2A_2(1\mathrm{cos}\theta 2\mathrm{cos}^2\theta \pm \varphi _{l,r}\mathrm{sin}\theta \pm 4\varphi _{l,r}\mathrm{sin}\theta \mathrm{cos}\theta )\\ +A_2^2(1+\mathrm{cos}\theta \varphi _{l,r}\mathrm{sin}\theta )]\\ (a_{q_3}^{l,r})^2=\frac{\alpha ^2+\beta ^2}{1+A_3^2}[13\varphi _{l,r}\mathrm{sin}\theta +3\mathrm{cos}\theta 4\mathrm{cos}^3\theta \pm 12\varphi _{l,r}\mathrm{cos}^2\theta sin\theta \\ +2A_3(1+\mathrm{cos}\theta 2\mathrm{cos}^2\theta \varphi _{l,r}\mathrm{sin}\theta \pm 4\varphi _{l,r}\mathrm{sin}\theta \mathrm{cos}\theta )\\ +A_3^2(1\mathrm{cos}\theta \pm \varphi _{l,r}\mathrm{sin}\theta )]\\ (a_{q_4}^{l,r})^2=\frac{\alpha ^2+\beta ^2}{1+A_4^2}[13\varphi _{l,r}\mathrm{sin}\theta +3\mathrm{cos}\theta 4\mathrm{cos}^3\theta \pm 12\varphi _{l,r}\mathrm{cos}^2\theta sin\theta \\ +2A_4(1+\mathrm{cos}\theta 2\mathrm{cos}^2\theta \varphi _{l,r}\mathrm{sin}\theta \pm 4\varphi _{l,r}\mathrm{sin}\theta \mathrm{cos}\theta )\\ +A_4^2(1\mathrm{cos}\theta \pm \varphi _{l,r}\mathrm{sin}\theta )].\end{array}$$
(20)
Letโs substitute eq. (19) into eq. (18), we get
$$\ddot{q}_\eta ^{l,r}+\omega _\eta ^2q_\eta ^{l,r}=(a_{q_\eta }^{l,r})\frac{eE_0}{m}\mathrm{cos}(\omega t+\sigma _{q_\eta }^{l,r})$$
(21)
and the solution of these equations is
$$q_\eta =(a_{q_\eta }^{l,r})\frac{eE_0}{m}\frac{\mathrm{cos}(\omega t+\sigma _{q_\eta }^{l,r})}{\omega _\eta ^2\omega ^2}.$$
(22)
The propagating light wave induces the dipole moments $`d_{q_\eta }^{l,r}`$ which are
$$d_{q_\eta }^{l,r}=q_\eta (a_{q_\eta }^{l,r})f_{q_\eta }e$$
(23)
or using (22)
$$d_{q_\eta }^{l,r}=(a_{q_\eta }^{l,r})^2\frac{f_{q_\eta }e^2E_0}{m}\frac{\mathrm{cos}(\omega t+\sigma _{q_\eta }^{l,r})}{\omega _\eta ^2\omega ^2},$$
(24)
where $`f_{q_\eta }`$ are the oscillator strengths in the normal modes of vibrations. The mean polarizability per volume unit is
$$\chi _{q_\eta }^{l,r}=\frac{N^{}d_{q_\eta }^{l,r}}{E_0\mathrm{cos}(\omega t+\sigma _{q_\eta }^{l,r})}=\frac{Nd_{q_\eta }^{l,r}}{3E_0\mathrm{cos}(\omega t+\sigma _{q_\eta }^{l,r})},$$
(25)
where $`N^{}`$ is the number of compound oscilators in the volume unit. We see that in our case $`N^{}=N/3`$ where $`N`$ is the number of isolated oscillators.
For the refractive indices of the crystals we have the Drude-Sellmaier dispersion relation
$$n_{l,r}^21=4\pi \underset{\eta =1}{\overset{4}{}}\chi _{q_\eta }^{l,r}=\frac{4\pi Ne^2}{3m}\underset{\eta =1}{\overset{4}{}}(a_{q_\eta }^{l,r})^2\frac{f_{q_\eta }}{\omega _\eta ^2\omega ^2}.$$
(26)
From eq. (26) we are able to solve the relation
$$n_l^2n_r^2=\frac{4\pi Ne^2}{3m}\underset{\eta =1}{\overset{4}{}}\frac{[(a_{q_\eta }^l)^2(a_{q_\eta }^r)^2]f_{q_\eta }}{\omega _\eta ^2\omega ^2}$$
(27)
and using (20) we can calculate for the expression in square brackets
$$\begin{array}{c}(a_{q_1}^l)^2(a_{q_1}^r)^2=\frac{(\alpha ^2+\beta ^2)(\varphi _l+\varphi _r)}{1+A_1^2}\mathrm{sin}\theta (312\mathrm{cos}^2\theta \\ 2A_18A_1\mathrm{cos}\theta A_1^2),\\ (a_{q_2}^l)^2(a_{q_2}^r)^2=\frac{(\alpha ^2+\beta ^2)(\varphi _l+\varphi _r)}{1+A_2^2}\mathrm{sin}\theta (312\mathrm{cos}^2\theta \\ 2A_28A_2\mathrm{cos}\theta A_2^2),\\ (a_{q_3}^l)^2(a_{q_3}^r)^2=\frac{(\alpha ^2+\beta ^2)(\varphi _l+\varphi _r)}{1+A_3^2}\mathrm{sin}\theta (3+12\mathrm{cos}^2\theta \\ 2A_3+8A_3\mathrm{cos}\theta +A_3^2),\\ (a_{q_4}^l)^2(a_{q_4}^r)^2=\frac{(\alpha ^2+\beta ^2)(\varphi _l+\varphi _r)}{1+A_4^2}\mathrm{sin}\theta (3+12\mathrm{cos}^2\theta \\ 2A_4+8A_4\mathrm{cos}\theta +A_4^2).\end{array}$$
(28)
In eq. (28) we can write $`(\varphi _l+\varphi _r)=\frac{\omega d(n_l+n_r)}{c}`$ and $`n_l^2n_r^2=(n_ln_r)(n_l+n_r)`$. Using the well known formula for the ORD $`\rho (\omega )=\frac{\omega }{2c}(n_ln_r)`$ we obtain
$`\rho (\omega )`$ $`=`$ $`{\displaystyle \frac{2\pi Nde^2}{3c^2}}\omega ^2(\alpha ^2+\beta ^2)\mathrm{sin}\theta `$ (29)
$`\times [{\displaystyle \frac{(312\mathrm{cos}^2\theta 2A_18A_1\mathrm{cos}\theta A_1^2)f_{q_1}}{(1+A_1^2)(\omega _1^2\omega ^2)}}`$
$`+{\displaystyle \frac{(312\mathrm{cos}^2\theta 2A_28A_2\mathrm{cos}\theta A_2^2)f_{q_2}}{(1+A_2^2)(\omega _2^2\omega ^2)}}`$
$`+{\displaystyle \frac{(3+12\mathrm{cos}^2\theta 2A_3+8A_3\mathrm{cos}\theta +A_3^2)f_{q_3}}{(1+A_3^2)(\omega _3^2\omega ^2)}}`$
$`+{\displaystyle \frac{(3+12\mathrm{cos}^2\theta 2A_4+8A_3\mathrm{cos}\theta +A_4^2)f_{q_4}}{(1+A_4^2)(\omega _4^2\omega ^2)}}].`$
Using eqs. (17) we can rewrite the formula (29) to the form which contains the natural frequency of the oscillators only but we see that we would get the complicated expression. We will discuss the result (29) only in the approximations of the oscillator strengths and also only for the crystals with the space groups of symmmetry $`D_3^4`$ and $`D_3^6`$ for which $`\theta =120`$ deg and therefore $`\mathrm{cos}\theta =1/2`$.
### 3.2 Discussion
We will express as in the sections 1, 2 and 3 the result (29) in the linear harmonic oscillator approximation and in the Heitler - London approximation of the oscillator strengths. In all results we neglect all the small terms containing the expressions $`1/(\omega _0^2\omega ^2)`$ of the order higher than second. These terms contain with the exception of the great value of the denominator also the higher orders of the small coupling constants $`Q_1`$ and $`Q_2`$ in the numerators.
At first we will use in the discussion of the result (29) the Chandrasekhar linear harmonic oscillator approximation where we can calculate with equation $`f_{q_1}=f_{q_2}=f_{q_3}=f_{q_4}=f_0`$. In this case we obtain the ORD formula of the Chandrasekhar type
$$\rho (\omega )=\frac{4\pi Nde^2f_0(\alpha ^2+\beta ^2)\mathrm{sin}\theta \left(Q_1\frac{4}{3}Q_2\right)}{mc^2}\frac{\omega ^2}{(\omega _0^2\omega ^2)^2}.$$
(30)
The Heitler-London aproximation is in the case of four coupled oscillators given by the relation
$$\frac{f_{q_1}}{\omega _1}=\frac{f_{q_2}}{\omega _2}=\frac{f_{q_3}}{\omega _3}=\frac{f_{q_4}}{\omega _4}=\frac{f_0}{\omega _0}$$
(31)
and using eqs. (17) and taking again into account that $`Q_1`$ and $`Q_2`$ are small quantities we can simplify expressions for the oscillator strengths to the form
$$\begin{array}{c}f_{q_1}=f_0(1+\frac{Q_1+\sqrt{5Q_1^2+8Q_1Q_2+4Q_2^2}}{4\omega _0^2}),\\ f_{q_2}=f_0(1+\frac{Q_1\sqrt{5Q_1^2+8Q_1Q_2+4Q_2^2}}{4\omega _0^2}),\\ f_{q_3}=f_0(1\frac{Q_1\sqrt{5Q_1^28Q_1Q_2+4Q_2^2}}{4\omega _0^2}),\\ f_{q_4}=f_0(1\frac{Q_1+\sqrt{5Q_1^28Q_1Q_2+4Q_2^2}}{4\omega _0^2}).\end{array}$$
(32)
Now we substitute these relations into eq. (29) and we obtain
$`\rho (\omega )`$ $`=`$ $`{\displaystyle \frac{2\pi Nde^2f_0}{3mc^2}}{\displaystyle \frac{\omega ^2}{\omega _0^2}}(\alpha ^2+\beta ^2)\mathrm{sin}\theta `$ (33)
$`\times \left[{\displaystyle \frac{3Q_14Q_2}{\omega _0^2\omega ^2}}+{\displaystyle \frac{2(3Q_1\omega _0^24Q_2\omega _0^22Q_1^2+2Q_1Q_2)}{(\omega _0^2\omega ^2)^2}}\right]`$
and after neglecting the terms $`Q_1^2`$ and $`Q_1Q_2`$ and adding the terms in the square brackets we have the formula of the Agranovich type
$$\rho (\omega )=\frac{2\pi Nde^2f_0(\alpha ^2+\beta ^2)\mathrm{sin}\theta (Q_1\frac{4}{3}Q_2)}{mc^2}\frac{\omega ^2(\omega _0^2+\omega ^2)}{\omega _0^2(\omega _0^2\omega ^2)^2}.$$
(34)
We see that applying the linear harmonic oscillator aproximation and the Heitler-London aproximation to the general ORD formula (29) of four coupled oscillator model we obtained for the crystals belonging to the space groups of symmetry $`D_3^4`$ and $`D_3^6`$ the Chandrasekhar (30) and the Agranovich (34) formulae again. Both formulae contain the same expression $`Q_1\frac{4}{3}Q_2`$. Letโs also note the results of solving three coupled oscillator model published in where these formulae contain the expression $`Q_1Q_2`$.
We obtain the similar results also in the five, six etc. oscillator models. In the linear harmonic oscillator and in the Heitler - London approximation we get again the Chandrasekhar and the Agranovich formulae which are different only in the constant containing the coupling constants $`Q_1`$ and $`Q_2`$. This constant has for the five oscillator model the form $`Q_1\frac{3}{2}Q_2`$, in the six oscillator model the form $`Q_1\frac{8}{5}Q_2`$ etc.
These results give the possibility to suppose the form of the ORD formulae for general $`\overline{N}`$ coupled oscillator model. We obtain the Chandrasekhar formula in the form
$$\rho (\omega )=\frac{4\pi Nde^2f_0(\alpha ^2+\beta ^2)\mathrm{sin}\theta \left(Q_1\frac{2(\overline{N}2)}{\overline{N}1}Q_2\right)}{mc^2}\frac{\omega ^2}{(\omega _0^2\omega ^2)^2},$$
(35)
where $`\overline{N}1`$ is the number of coupling between neighbouring oscillators and $`\overline{N}2`$ the number of coupling between odd and even oscillators in the model. It means that for the practical case $`\overline{N}\mathrm{}`$ we have
$$\rho (\omega )=\frac{4\pi Nde^2(\alpha ^2+\beta ^2)\mathrm{sin}\theta \left(Q_12Q_2\right)}{mc^2}\frac{\omega ^2}{(\omega _0^2\omega ^2)^2}.$$
(36)
The same results we can write for the Agranovich formula:
$$\rho (\omega )=\frac{2\pi Nde^2f_0(\alpha ^2+\beta ^2)\mathrm{sin}\theta (Q_1\frac{2(\overline{N}2)}{\overline{N}1}Q_2)}{mc^2}\frac{\omega ^2(\omega _0^2+\omega ^2)}{\omega _0^2(\omega _0^2\omega ^2)^2}$$
(37)
and for $`\overline{N}\mathrm{}`$
$$\rho (\omega )=\frac{2\pi Nde^2f_0(\alpha ^2+\beta ^2)\mathrm{sin}\theta (Q_12Q_2)}{mc^2}\frac{\omega ^2(\omega _0^2+\omega ^2)}{\omega _0^2(\omega _0^2\omega ^2)^2}.$$
(38)
We see that the expression $`Q_12Q_2`$ is the common limit of the coefficients containing the coupling constants in the final ORD formulae in the case $`\overline{N}\mathrm{}`$.
## 4 Conclusions
In this paper we have proved that with the Chandrasekhar ORD two coupled oscillator model we can include all couplings between adjacent oscillators on the helix that is given by the crystal structure. We have solved that the influence of the couplings between adjacent oscillators in the cases when the second single oscillator in the first compound oscillator is the first oscillator in the second compound oscillator we can hold as aditive (the mistake in this conclusion is neglectable). For the including of all couplings between adjacent oscillators is sufficient to take in the Chandrasekhar model that the number of single oscillators in the volume unit is the same as the number of the couplings between adjacent oscillators. The same results were proved for the models of three and four coupled oscillators - the parts of the ORD results that contain the coupling constant between adjacent oscillators $`Q_1`$ are identical in all models.
From the crystal structure it follows that we cannot neglect the couplings between even and odd oscillators on the helix. We can assume for the real crystal that the value of the coupling constant between even and odd oscillators $`Q_2`$ is comparable with the value of the coupling constant between adjacent oscillators $`Q_1`$. Further we have proved that in the crystals with the space groups of symmetry $`D_3^4`$ and $`D_3^6`$ the couplings between even and odd oscillators have the opposite effect on the ORD than the couplings between adjacent oscillators.
The first coupled oscillator model in which we can include the coupling between even or odd oscillators is the three oscillator model. But using the previous conclusion this model can include all couplings between adjacent oscillators and only a half of couplings between even and odd oscillators on the helix. Besides we cannot assume these couplings as aditive because they function between single oscillators inside of the compound oscillators - for example in the three oscillator model the second oscillator in the compound oscillator is in the fact coupled also with the second oscillator in the adjacent compound oscillator (becouse in the used model the last single oscillator in any compound oscillator is the first single oscillator in the adjacent compound oscillator).
From this reason we have further solved the models of four, five etc. coupled oscillators. In these models the number of the included couplings between even and odd oscillators increases and the results can be generalized. By this way we have proved that the couplings between even and odd oscillators also have the aditive effects. The ORD results of the three oscillators model after the approximations of the oscillator strengths of the normal modes of vibrations contain the coefficient $`Q_1Q_2`$ (for the crystals with space groups of symmetry $`D_3^4`$ and $`D_3^6`$). In the limit of $`\overline{N}`$ coupled oscillators for $`\overline{N}\mathrm{}`$ it was derived the coefficient $`Q_12Q_2`$. This also means that the relative influence of the couplings between even and odd oscillators on the ORD is twofold in comparison with the couplings between adjacent oscillators.
We can formally obtain the generalized ORD results of the $`\overline{N}`$ coupled oscillators model after the approximations of the oscillator strengths using the three oscillators model in which we denote the value of coupling between the first and the third oscillators as $`2Q_2`$. We can say that this conclusion is practically acceptable only after using the results of the three oscillator model in the numerical approximations of the experimental ORD data for example of $`\alpha `$-quartz or tellur.
|
warning/0006/nucl-th0006059.html
|
ar5iv
|
text
|
# Projected shell model study for the yrast-band structure of the proton-rich mass-80 nuclei
## 1 Introduction
The study of low- and high-spin phenomena in the proton-rich mass-80 nuclei have attracted considerable interest in recent years. This has been motivated by the increasing power of experimental facilities and improved theoretical descriptions, as well as by the astrophysical requirement in understanding the structure of these unstable nuclei. In comparison to the rare-earth region where the change in nuclear structure properties is quite smooth with respect to particle number, the structure of the proton-rich mass-80 nuclei shows considerable variations when going from one nucleus to another. This is mainly due to the fact that the available shell model configuration space in the mass-80 region is much smaller than in the rare-earth region. The low single particle level density implies that a drastic change near the Fermi surfaces can occur among neighboring nuclei. Another fact is that in these medium mass proton-rich nuclei, neutrons and protons occupy the same single particle orbits. As the nucleus rotates, pair alignments of neutrons and protons compete with each other and in certain circumstances they can align simultaneously.
Thanks to the new experimental facilities, in particular, to the newly constructed detector arrays, the domain of nuclides accessible for spectroscopic studies has increased drastically during the past decade. For example, some extensive measurements of the transition quadrupole-moments, extracted from the level lifetimes and excitation energy of 2<sup>+</sup> state, have been carried out for Kr-, Sr- and Zr-isotopes. These measurements have revealed large variations in nuclear structure of these isotopes with respect to particle number and angular momentum. It has been shown that alignment of proton- and neutron-pairs at higher angular momenta can change the nuclear shape from prolate to triaxial and to oblate.
A systematic description for all these observations poses a great challenge to theoretical models. The early mean-field approaches were devoted to a general study of the structure in the mass-80 region. Besides the work using the Woods-Saxon approach , there were studies using the Nilsson model and the Skyrme Hartree-Fock + BCS theory . More recently, microscopic calculations were performed using the Excited VAMPIR approach . However, this approach is numerically quite involved and is usually employed to study some quantities for selected nuclei. The large-scale spherical shell model diagonalization calculations have been recently successful in describing the $`pf`$ shell nuclei, but the configuration space required for studying the well-deformed mass-80 nuclei is far beyond what the modern computers can handle.
Recently, the projected shell model (PSM) has become quite popular to study the structure of deformed nuclei. The advantage in this method is that the numerical requirement is minimal and, therefore, it is possible to perform a systematic study for a group of nuclei in a reasonable time frame. The PSM approach is based on the diagonalization in the angular-momentum projected basis from the deformed Nilsson states. A systematic study of the rare-earth nuclei has been carried out and the agreement between the PSM results and experimental data has been found to be quite good. Very recently, the PSM approach has also been used to study the high-spin properties of <sup>74</sup>Se , which lies in the mass-80 region.
The purpose of the present work is to perform a systematic PSM study of the low- and high-spin properties for the proton-rich, even-even Kr-, Sr- and Zr-isotopes. The physical quantities to be described are energy spectrum, transition quadrupole moment and gyromagnetic factor. In addition, to compare the avaliable data with theory in a systematic way, we also make predictions for the structure of the $`N=Z`$ nuclei which could be tested by future experiments with radioactive ion beams. We shall begin our discussion with an outline of the PSM in section II. The results of calculations and comparisons with experimental data are presented in section III. Finally, the conclusions are given in section IV.
## 2 The Projected shell model
In this section, we shall briefly outline the basic philosophy of the PSM. For more details about the model, the reader is referred to the review article . The PSM is based on the spherical shell model concept. It differs from the conventional shell model in that the PSM uses the angular momentum projected states as the basis for the diagonalization of the shell model Hamiltonian. What one gains by starting from a deformed basis is not only that shell model calculations for heavy nuclei become feasible but also physical interpretation for the complex systems becomes easier and clearer.
The wave function in the PSM is given by
$$|\sigma ,IM=\underset{K,\kappa }{}f_\kappa ^\sigma \widehat{P}_{MK}^I|\varphi _\kappa .$$
(1)
The index $`\sigma `$ labels the states with same angular momentum and $`\kappa `$ the basis states. $`\widehat{P}_{MK}^I`$ is angular momentum projection operator and $`f_\kappa ^\sigma `$ are the weights of the basis states $`\kappa `$.
We have assumed axial-symmetry for the basis states and the intrinsic states are, therefore, the eigenstates of the $`K`$-quantum number. For calculations of an even-even system, the following four kinds of basis states $`|\varphi _\kappa `$ are considered: the quasiparticle (qp) vacuum $`|0`$, two-quasineutron states $`a_{\nu _1}^{}a_{\nu _2}^{}|0`$, two-quasiproton states $`a_{\pi _1}^{}a_{\pi _2}^{}|0`$, and two-quasineutron plus two-quasiproton (or 4-qp) states $`a_{\nu _1}^{}a_{\nu _2}^{}a_{\pi _1}^{}a_{\pi _2}^{}|0`$. The projected vacuum $`|0`$, for instance, is the ground-state band (g-band) of an even-even nucleus.
The weight factors, $`f_\kappa ^\sigma `$ in Eq. (1), are determined by diagonalization of the shell model Hamiltonian in the space spanned by the projected basis states given above. This leads to the eigenvalue equation
$$\underset{\kappa ^{}}{}(H_{\kappa \kappa ^{}}E_\sigma N_{\kappa \kappa ^{}})f_\kappa ^{}^\sigma =0,$$
(2)
and the normalization is chosen such that
$$\underset{\kappa \kappa ^{}}{}f_\kappa ^\sigma N_{\kappa \kappa ^{}}f_\kappa ^{}^\sigma ^{}=\delta _{\sigma \sigma ^{}},$$
(3)
where the Hamiltonian and norm matrix-elements are given by
$`H_{\kappa \kappa ^{}}`$ $`=`$ $`\varphi _\kappa |\widehat{H}\widehat{P}_{K_\kappa K_\kappa ^{}^{}}^I|\varphi _\kappa ^{},`$ (4)
$`N_{\kappa \kappa ^{}}`$ $`=`$ $`\varphi _\kappa |\widehat{P}_{K_\kappa K_\kappa ^{}^{}}^I|\varphi _\kappa ^{}.`$ (5)
In the numerical calculations, we have used the standard quadrupole-quadrupole plus (monopole and quadrupole) pairing force, i.e.
$$\widehat{H}=\widehat{H_0}\frac{1}{2}\chi \underset{\mu }{}\widehat{Q}_{\mu }^{}{}_{}{}^{}\widehat{Q}_\mu G_M\widehat{P}^{}\widehat{P}G_Q\underset{\mu }{}\widehat{P}_{\mu }^{}{}_{}{}^{}\widehat{P}_\mu ,$$
(6)
where $`\widehat{H_0}`$ is the spherical single-particle Hamiltonian. The strength of the quadrupole force $`\chi `$ is adjusted such that the known quadrupole deformation parameter $`ฯต_2`$ is obtained. This condition results from the mean-field approximation of the quadrupole-quadrupole interaction of the Hamiltonian in Eq. (6). The monopole pairing force constants $`G_M`$ are adjusted to give the known energy gaps. For all the calculations in this paper, we have used
$`G_M^\nu `$ $`=`$ $`\left[20.2516.20{\displaystyle \frac{NZ}{A}}\right]A^1,`$
$`G_M^\pi `$ $`=`$ $`20.25A^1.`$ (7)
The strength parameter $`G_Q`$ for quadrupole pairing is assumed to be proportional to $`G_M`$. It has been shown that the band crossing spins vary with the magnitude of the quadrupole pairing force . The ratio $`G_Q/G_M`$ has therefore been slightly adjusted around 0.20 to reproduce the observed band crossings.
Electromagnetic transitions can give important information on the nuclear structure and provide a stringent test of a particular model. In the present work, we have calculated the electromagnetic properties using the PSM approach. The reduced transition probabilities $`B(EL)`$ from the initial state $`(\sigma _i,I_i)`$ to the final state $`(\sigma _f,I_f)`$ are given by
$$B(EL,I_iI_f)=\frac{e^2}{(2I_i+1)}|\sigma _f,I_f||\widehat{Q}_L||\sigma _i,I_i|^2,$$
(8)
where the reduced matrix-element is given by
$`\sigma _f,I_f\widehat{Q}_L\sigma _i,I_i`$ $`=`$ $`{\displaystyle \underset{\kappa _i,\kappa _f}{}}f_{\kappa _i}^{\sigma _i}f_{\kappa _f}^{\sigma _f}{\displaystyle \underset{M_i,M_f,M}{}}()^{I_fM_f}\left(\begin{array}{ccc}I_f& L& I_i\\ M_f& M& M_i\end{array}\right)`$
$`\times \varphi _{\kappa _f}|\widehat{P}_{}^{I_f}{}_{K_{\kappa _f}M_f}{}^{}\widehat{Q}_{LM}\widehat{P}_{K_{\kappa _i}M_i}^{I_i}|\varphi _{\kappa _i}`$
$`=`$ $`2{\displaystyle \underset{\kappa _i,\kappa _f}{}}f_{\kappa _i}^{\sigma _i}f_{\kappa _f}^{\sigma _f}{\displaystyle \underset{M^{},M^{\prime \prime }}{}}()^{I_fK_\kappa f}(2I_f+1)^1\left(\begin{array}{ccc}I_f& L& I_i\\ K_{\kappa _f}& M^{}& M^{\prime \prime }\end{array}\right)`$ (12)
$`\times {\displaystyle }d\mathrm{\Omega }D_{M^{\prime \prime }K_{\kappa _i}}(\mathrm{\Omega })\varphi _{\kappa _f}|\widehat{Q}_{LM^{}}\widehat{R}(\mathrm{\Omega })|\varphi _{\kappa _i}.`$ (13)
The transition quadrupole moment $`Q_t(I)`$ is related to the $`B(E2)`$ transition probability through
$$Q_t(I)=\left(\frac{16\pi }{5}\frac{B(E2,II2)}{IK20|I2K}\right)^{1/2}.$$
(14)
In the calculations, we have used the effective charges of 1.5e for protons and 0.5e for neutrons, which are the same as in the previous PSM calculations .
Variation of $`Q_t`$ with spin $`I`$ can provide information on shape evolution in nuclei. In the case of a rigid rotor, the $`Q_t`$ curve as a function of $`I`$ is a straight line. Experimentally, one finds a deviation from the rigid body behavior for most of the nuclei in this mass region. One expects more or less a constant value in $`Q_t`$ up to the first band crossing. At the band crossing, one often sees a dip in $`Q_t`$ value due to a small overlap between the wave functions of the initial and the final states involved. There can be a change in $`Q_t`$ values after the first band crossing, which indicates a shape change induced by quasiparticle alignment.
The other important electromagnetic quantity, which can give crucial information about the level occupancy and thus is an direct indication of the nature of alignment, is the gyromagnetic factor (g-factor). The g-factors $`g(\sigma ,I),g_p(\sigma ,I)`$ and $`g_n(\sigma ,I)`$ are defined by
$$g(\sigma ,I)=\frac{\mu (\sigma ,I)}{\mu _NI}=g_\pi (\sigma ,I)+g_\nu (\sigma ,I),$$
(15)
with $`g_\tau (\sigma ,I),\tau =\pi ,\nu `$ given by
$`g_\tau (\sigma ,I)`$ $`=`$ $`{\displaystyle \frac{1}{\mu _N[I(I+1)]^{1/2}}}`$
$`\times \left[g_l^\tau <\sigma ,I||\widehat{J}^\tau ||\sigma ,I>+(g_s^\tau g_l^\tau )\sigma ,I\widehat{S}^\tau \sigma ,I\right].`$ (16)
In our calculations, the following standard values of $`g_l`$ and $`g_s`$ have been taken: $`g_l^\pi =1,g_l^\nu =0,g_s^\pi =5.586\times 0.75`$ and $`g_s^\nu =3.826\times 0.75`$. Unfortunately, very few experiments have been done to measure the g-factors in this mass region. The only experiment which has recently been performed is for <sup>84</sup>Zr . In the present paper, we have calculated g-factors for this nucleus and compared with the data.
## 3 Results and Discussions
The PSM calculations proceed in two steps. In the first step, an optimum set of deformed basis is constructed from the standard Nilsson potential. The Nilsson parameters are taken from Ref. and the calculations are performed by considering three major shells (N = 2, 3 and 4) for both protons and neutrons. The basis deformation $`ฯต_2`$ used for each nucleus is taken either from experiment if measurement has been done or from the theoretical value of the total routhian surface (TRS) calculations. The intrinsic states within an energy window of 3.5 MeV around the Fermi surface are considered. This gives rise to the size of the basis space, $`|\varphi _\kappa `$ in Eq. (1), of the order of 40. In the second step, these basis states are projected to good angular-momentum states, and the projected basis is then used to diagonalize the shell model Hamiltonian. The detailed calculations have been carried out for Kr-, Sr- and Zr-isotopes and the results are discussed in the following subsections. The band diagram , which gives the projected energies for the configurations close to the Fermi surface is shown in Figs. 1,2 and 3 to explain the underlying physics. It should be noted that only a few of the most important configurations are plotted in the band diagrams, although many more are included in the calculations.
### 3.1 Kr isotopes
The proton-rich Kr-isotopes depict a variety of coexistence of prolate and oblate shapes in the low-spin region. For these nuclei, the low-lying states have an oblate deformation and the states above $`I^\pi =4^+`$ are associated with a prolate shape.
The first experimental information on transitions in <sup>72</sup>Kr was later confirmed in Refs. . Recently, the level scheme has been extended up to $`I^\pi =16^+`$ . The experimental value of the transition strength is available only for $`I^\pi =8^+`$. Following the indication in Ref. that the yrast band at higher spins is prolate in nature, we have calculated the <sup>72</sup>Kr with a prolate basis deformation of $`ฯต_2=0.36`$. As shown in the band diagram in Fig. 1, the g-band is crossed by a pair of proton 2-qp bands and a neutron 2-qp band at about $`I=10`$. This gives rise to a simultaneous alignment of neutron and proton pairs. The measured energies and moment of inertia are compared with the calculated values in Figs. 4 and 5. The calculated $`Q_t`$ values are shown in Fig. 6. The only data point available is in good agreement with our calculations.
Recent lifetime measurements in <sup>74</sup>Kr up to $`I^\pi =18^+`$ indicate a pronounced shape change induced by quasiparticle alignments. The $`[440]\frac{1}{2}`$, $`[431]\frac{3}{2}`$, $`[422]\frac{5}{2}`$ neutron orbitals and the $`[440]\frac{1}{2}`$, $`[431]\frac{3}{2}`$, $`[422]\frac{5}{2}`$ proton orbitals play a major role in determining the nuclear shape at the high-spin region. In order to describe the high-spin states properly, we have used the prolate deformation of $`ฯต_2`$ = 0.37 in the PSM calculations for <sup>74</sup>Kr. The band diagram is shown in Fig. 1. The neutron 2-qp state $`\nu g_{9/2}[3/2,5/2]`$ and two proton 2-qp states $`\pi g_{9/2}[1/2,3/2]`$ and $`\pi g_{9/2}[3/2,5/2]`$ cross the g-band around $`I=14`$. This corresponds to the alignment of neutron and proton pairs as observed in the experiment. The yrast band, consisting of the lowest energies after diagonalization at each spin, is plotted in Fig. 4 to compare with the available experimental data. These values are also displayed in Fig. 5 in a sensitive plot of moment of inertia as a function of square of the rotational frequency. For <sup>74</sup>Kr, there is a very good agreement between the theory and experiment in both plots, except for the lowest spin states. In Fig. 6, the $`Q_t`$ values from the theory and experiment are compared. The sudden fall in the $`Q_t`$ at $`I=14`$, which is described correctly by our calculations, is associated with the first band crossing (see Fig. 1). Due to the occupation of the fully aligned orbitals after this band crossing, the $`Q_t`$ value becomes smaller for the higher spin states. Above $`I=18`$, maximum contribution to the yrast states comes from the configuration based on the 4-qp state $`\nu g_{9/2}[3/2,5/2]`$ \+ $`\pi g_{9/2}[1/2,3/2]`$.
The energy levels and their lifetimes in <sup>76</sup>Kr have been measured up to $`I^\pi =22^+`$ along the yrast band . Lifetime measurements indicate a large deformation in this nucleus. A simultaneous alignment of proton and neutron pairs is observed. For this nucleus, the calculations are performed for a prolate deformation of $`ฯต_2`$ = 0.36. The band diagram of <sup>76</sup>Kr is shown in Fig. 1. The g-band is crossed by the neutron 2-qp band $`\nu g_{9/2}[3/2,5/2]`$ and two proton 2-qp bands $`\pi g_{9/2}[1/2,3/2]`$ and $`\pi g_{9/2}[3/2,5/2]`$ between $`I=1214`$. This corresponds to the simultaneous alignments of neutron and proton pairs. At higher spins, the wave function of the yrast states receives contribution from the 4-qp configurations. Though the major component of the wave function for the higher-spin states comes from the 4-qp state based on $`\nu g_{9/2}[3/2,5/2]+\pi g_{9/2}[1/2,3/2]`$ as shown in the plot, we find some amount of $`K`$-mixing with other 4-qp states leading to a non-axial shape. The comparisions of the calculated energies and $`Q_t`$ values with experimental data show good agreement as shown in Figs. 4, 5 and 6. As in the <sup>74</sup>Kr case, the drop in $`Q_t`$ value at spin $`I=14`$ is obtained due to the occupation of the aligned states.
Although good agreement between theory and experiment for the <sup>72</sup>Kr spectrum is obtained in Fig. 4, clear discrepancy can be found in the more sensitive plot of Fig. 5. Besides the problem in the description of the moment of inertia for the lowest spin states as in the <sup>74</sup>Kr and <sup>76</sup>Kr cases, the current calculation can not reproduce the fine details around the backbending region. For this $`N=Z`$ nucleus, there has been an open question of whether the proton-neutron pair correlation plays a role in the structure discussions. It has been shown that with the renormalized pairing interactions within the like-nucleons in an effective Hamiltoinan, one can account for the $`T=1`$ part of the proton-neutron pairing . However, whether the renormalization is sufficient for the complex region that exhibits the phenomenon of band crossings, in particular when both neutron and proton pair alignments occur at the same time is an interesting question to be investigated.
### 3.2 Sr isotopes
Light strontium isotopes are known to be among the well-deformed nuclei in the mass-80 region. In contrast to the Kr isotopes, data for Sr, including the most recent ones , do not suggest a backbending or upbending in moment of inertia (see Fig. 5). For <sup>76</sup>Sr, the only experimental work was reported in Ref. , where the energy of the first excited state was identified. Because of shell gap at $`N=38`$, this nucleus is found to be highly deformed. We have calculated this nucleus with the basis deformation of $`ฯต_2`$ = 0.36. It is found in Fig. 2 that the proton 2-qp band $`\pi g_{9/2}[3/2,5/2]`$ and the neutron 2-qp band $`\nu g_{9/2}[3/2,5/2]`$ nearly coincide with each other for the entire low-spin region, and cross the g-band at $`I=14`$. Just above $`I=16`$, the 4-qp band based on the configuration $`\nu g_{9/2}[3/2,5/2]+\pi g_{9/2}[3/2,5/2]`$ crosses the 2-qp bands and becomes the lowest one for higher spins. However, both of the first and the second band crossings are gentle with very small crossing angles. Therefore, the structural change along the yrast band is gradual, and the yrast band obtained as the lowest band after band mixing seems to be unperturbed. This is true also for the isotopes <sup>78</sup>Sr and <sup>80</sup>Sr as we shall see below. Thus, although band crossings occur in both of the Kr and Sr isotopes, smaller crossing angle results in a smooth change in the moments of inertia in Sr, and this picture is very different from what we have seen in the Kr band diagrams in Fig. 1, where sharper band crossings lead to sudden structure changes.
Ref. reported the positive parity yrast cascade in <sup>78</sup>Sr up to $`I^\pi =10^+`$. Their measurement of the lifetimes of the first two excited states indicate the $`E2`$ transition strengths of more than 100 Weisskopf units. Refs. extended this sequence up to $`I^\pi =22^+`$ . A broad band crossing was observed around $`I=12`$ and two bands in <sup>78</sup>Sr seem to interact strongly over a wide range of spin states. To study these effects, the band energies are calculated for <sup>78</sup>Sr, assuming a prolate deformation $`ฯต_2`$ = 0.36. As shown in the band diagram of <sup>78</sup>Sr in Fig. 2, the band based on the proton 2-qp state occupying $`[431]\frac{3}{2}`$ and $`[422]\frac{5}{2}`$ orbitals crosses the g-band at $`I=12`$ at a very small angle. This implies a smooth structure change in the yrast band and explains the gradual alignment behavior as observed in the experiment. In fact, the neutron 2-qp band based on $`[431]\frac{3}{2}`$ and $`[422]\frac{5}{2}`$ orbitals contribute to the yrast levels between $`I=14`$ to 18. The calculated energies and moment of inertia are compared with the experimental data in Figs. 4 and 5. In Fig. 6, the calculated $`Q_t`$ values show a smooth behavior with a slight decreasing trend for the entire spin region, which is in contrast to the sudden drop in the $`Q_t`$ curve at $`I=14`$ in its isotone <sup>74</sup>Kr. The current experimental data exist only for the lowest two states. To test our prediction, the measurements for higher spins are desirable.
In a recent experiment , <sup>80</sup>Sr has been studied up to $`I^\pi =22^+`$ along the yrast band. To observe the shape evolution along the yrast band, the lifetime of levels up to $`I=14^+`$ has been measured. At low spin, a prolate shape with $`\beta _20.35`$ has been reported . Taking $`ฯต_2=0.34`$, the band diagram of <sup>80</sup>Sr is calculated and displayed in Fig. 2. A proton 2-qp band based on $`\pi g_{9/2}[3/2,5/2]`$ crosses the g-band between $`I=10`$ and 12 and this corresponds to the first proton pair alignment. At higher spin values, the yrast states get contribution from a large number of 4-qp configurations with different $`K`$-values (These bands are not shown in Fig. 2). The calculated spectrum is compared with data in Figs. 4 and 5. The agreement is satisfactory. In particular, the smooth evolution of moments of inertia is reproduced in Fig. 5. Two sets of measured $`Q_t`$ values are compared with our calculations. The pronounced fluctuations in the $`Q_t`$ values from the early experiment have been removed by the new measurement in Ref. . However, our calculation does not reproduce the newly measured $`Q_t`$ at $`I=14`$.
### 3.3 Zr isotopes
The self-conjugate isotope <sup>80</sup>Zr was studied up to $`I^\pi =4^+`$ by the in-beam study . The excitation energy of the 2<sup>+</sup> state indicates a large deformation for this nucleus. In Fig. 3, the band diagrams of the proton-rich Zr isotopes are shown. The band diagram of <sup>80</sup>Zr is calculated with $`ฯต_2`$ = 0.36. For the positive parity band, the 3/2 and 5/2 quasiparticle states lie close to the Fermi levels of both neutrons and protons. These orbitals play an important role in determining the alignment properties in <sup>80</sup>Zr. Between $`I=12`$ and 14, the proton and neutron 2-qp bands cross the ground band at the same spin. These 2-qp bands consist of two $`g_{9/2}`$ quasiparticles in the 3/2 and 5/2 orbitals, coupled to $`K=1`$. After this band crossing, the yrast band gets maximum contribution from both of the proton 2-qp and neutron 2-qp states, as can be seen in Fig. 3. Immediately following the first band crossing, a 4-qp band crosses the 2-qp bands at $`I=16`$. In the plots of moment of inertia in Fig. 5, we see that the total effect of the first and second band crossings produces a smooth upbending. In the $`Q_t`$ plots in Fig. 6, a two-step drop in the $`Q_t`$ values is predicted, which is due to the two successive band crossings.
Let us now look at the $`N=Z+2`$ isotope of Zr. Recently, <sup>82</sup>Zr has been studied up to $`I^\pi =24^+`$ . Lifetime measurements up to $`I^\pi =14^+`$ along the yrast line indicate a smaller deformation $`ฯต_2`$ = 0.29 for this nuclei. The band diagram is shown in Fig. 3. Here, the 2-qp band consists of two $`g_{9/2}`$ quasi-protons in 3/2 and 5/2 orbitals crosses the g-band at $`I=10`$. Beyond that spin, this 2-qp proton configuration contributes maximum to the yrast states up to $`I=16`$, where a 4-qp band crosses the proton 2-qp band. This 4-qp configuration is made from two quasi-protons in 3/2 and 5/2 orbitals and two quasi-neutrons in 3/2 and 5/2 orbitals. These two separate band crossings were observed in the experiment . As can be seen from Figs. 4 and 5, a remarkable agreement between our calculation and experimental data are obtained. The observed complex behavior of the two-step upbending in moment of inertia in <sup>82</sup>Zr is correctly reproduced. As shown in Fig. 6, for <sup>82</sup>Zr, our theory predicts a small kink at $`I=10`$ in the $`Q_t`$ value corresponding to the band crossing of the proton 2-qp band. At present the accuracy of the experimental values of $`Q_t`$ is not enough to confirm this kink. Up to $`I=12`$, the measured transition quadrupole moments match quite well with the calculations. At spin $`I=16`$, the dip in the calculated $`Q_t`$ value is because of the second band crossing. A close look at the components of the wave functions of the yrast states, shows the mixing of different intrinsic states increases after the first band crossing around $`I=10`$. Thus this nucleus loses the axial symmetry at higher spin states.
For the $`N=Z+4`$ isotope of Zr, the deformation is less than the lighter ones . The band diagram shown in Fig. 3 is based on calculations with the basis deformation of $`ฯต_2=0.22`$. At least two proton 2-qp bands based on $`\pi g_{9/2}[3/2,5/2]`$ and $`\pi g_{9/2}[1/2,5/2]`$ configurations cross the g-band at $`I=8`$, corresponding to the proton alignment. After the first band crossing till $`I=14`$, these two 2-qp bands lie together and continue to interact with each other. Then one finds a neutron alignment due to the lowering of a 4-qp band between spin $`I=14`$ and 16. Starting from $`I=18`$, two 4-qp bands nearly coincide for the higher spin states. The calculated energies of the yrast band are compared with that of the adopted level scheme in Figs. 4 and 5. In Fig. 5, one can see that the interesting pattern of experimental moment of inertia in <sup>84</sup>Zr has been qualitatively reproduced although our calculation exaggerates the variations in the data. To see the shape evolution up to high spins, the experimental transition quadrupole moments are compared with the calculated values in Fig. 6. The measured $`Q_t`$ values for <sup>84</sup>Zr show different behavior from those of the neighboring even-even nuclei with an increasing trend after the second band crossing. Though the calculated values of $`Q_t`$ at higher spins are lower than the experimental data, the qualitative variation is well reproduced.
The only experiment that measured g-factors has recently been performed for <sup>84</sup>Zr . Though, there are big uncertainties, the three data points suggest the rigid rotor value $`Z/A`$. As we can see in the theoretical diagram given in Fig 7, g-factors for the low spin states in this nucleus are predicted to have a rigid body character with a nearly constant value up to $`I=6`$. However, a more than doubled value is predicted for $`I=8`$ with a decreasing trend thereafter. This pronounced jump in g-factor is obviously due to the proton 2-qp band crossings at $`I=8`$, as discussed above (see also Fig. 3). The wave functions of yrast states at $`I=8`$ to 14 are dominated by the proton 2-qp states, thus enhancing the g-factor values. Therefore, measurement extended to higher spins in this nucleus will be a strong test for our predictions. Similar g-factor increasing at the first band crossing should also appear in the neighboring nuclei, for example in <sup>82</sup>Zr, where proton 2-qp bands alone dominate the yrast states.
## 4 Conclusion
The study of the proton-rich mass-80 nuclei is not only interesting from the structure point of view, but it also has important implications in the nuclear astrophysical study . In this paper, we have for the first time performed a systematic study for the yrast bands of $`T_z`$ = 0, 1 and 2 isotopes of Kr, Sr and Zr within the framework of the projected shell model approach. We have employed the quadrupole plus monopole and quadrupole pairing force in the Hamiltonian, and the major shells with N = 2, 3 and 4 have been included for the configuration space.
The proton-rich mass-80 nuclei exhibit many phenomena that are quite unique to this mass region. The structure changes are quite pronounced among the neighboring nuclei and can be best seen in the sensitive plot of Fig. 5. For the Kr nuclei, a clear backbending is observed for all the three isotopes, while for the Sr isotopes no backbending is seen. However, the Zr nuclei exhibit both the first and the second upbends. The transition quadrupole moments show corresponding variations. For these variations, we have obtained an overall qualitative and in many cases quantitative description. These variations can be understood by the mixing of various configurations of the projected deformed Nilsson states. In particular, they are often related to band crossing phenomenon.
Despite the success mentined above, our calculations fail to reproduce states at very low spins, as can be clearly seen in the moment of inertia plot of Fig. 5 and the $`Q_t`$ plot of Fig. 6. The present model space which is constructed from the intrinsic states with a fixed axial deformation may be too crude for the spin region characterized by shape coexistence. To correctly describe the low-spin region, one needs to enrich the shell model basis. Introduction of triaxiality in the deformed basis combined with three dimensional angular momentum projection certainly spans a richer space. However, in order to study the high-spin states discussed in this paper, the current space of the triaxial PSM must be extended by including multi-quasiparticle states. The other possibility is to perform an investigation with the generator coordinate method which takes explicitly the shape evolution into account.
|
warning/0006/cs0006004.html
|
ar5iv
|
text
|
# A Note on โOptimal Static Load Balancing in Distributed Computer Systemsโ
## 1 Introduction
Recent years have been witness to an increasing use of distributed computing system. This may be attributed to two main factors: growth of the Internet, and low cost solution of end-user computing devices. Many business processes such as supply chain management are distributed due to the inherent nature of tasks involved with them. Besides, scale of economy is often possible due to the use of clusters of less powerful computers instead of a central computer of significantly high power. However, a distributed solution can yield the true advantage only if it is possible to distribute works evenly among the processors (terms processor, host and computer are used inter-changeably in this article). In other words, when load on the computers in a distributed environment has significant variance of workloads, high performance can be achieved by redistributing loads. The task of redistributing the loads on the computers is called load balancing.
Load can be characterized as jobs or tasks. A job indicates a complete and independent entity of work to be completed. However, a job may consist of number of sub-units called tasks, and for the job to be completed the comprising tasks need to often communicate among themselves. For simplification, in this article, we consider mutually independent tasks units as jobs. Jobs can be classified into two broad categories - dedicated and generic. Dedicated jobs can be processed only on specified hosts, while generic jobs can be processed on any host in the system. Once a generic job is submitted to the system, it can be processed at the place of its origin or can be transferred it to another processor. Irrespective of the processor chosen, once the job is started at a processor, it remains there till completion. Some of the most commonly used performance indicators of a load balancing algorithm are mean response time, throughput, variance of response time. In this article, we confine our analysis based on mean response time only.
Algorithms or heuristics for load balancing fall into two categories: static and dynamic. Static load balancing techniques do not depend on the state (characterized by workload etc.) of the processors. Simple static load balancing techniques such as join-the-shortest queue (SQ), minimum expected delay (MED) have been in use for over a decade. The SQ policy allocates an arriving task to the processor having the minimum load and, the MED policy allocates an arriving task to the processor having the minimum expected value of waiting time for currently scheduled tasks. Note that though these two policies may work similarly for homogenous processors, their behaviors differ for heterogeneous processors. Tantawi and Towsley have proposed an optimization model and load balancing algorithms to determine static optimal allocation of loads among the hosts connected in a broadcast network. Ross et al. considered a more general problem consisting of dedicated and generic jobs, and also dealt with scheduling decision at each host. The authors have noted that the problem is separable over local scheduling decisions, and suggested a solution procedure based on this finding.
Dynamic load balancing techniques, on the other hand, rely on present and past states of the processors. In dynamic load balancing, though an initial assignment of work to the processors is done through static load balancing, subsequent adjustment of work depends on the workload profile of the processors. Dynamic load balancing is expected to have higher overhead, but better performance than static load balancing. Most of the dynamic load balancing algorithms have two important components: Information exchange through which processors exchange their load information, and subsequent load exchange to redistribute loads. Depending on the characteristics of these components, there can be different types of dynamic load balancing ( , , ).
However, in this article we concentrate on static load balancing. We express some of the points of our disagreement with and provide alternatives. The rest of the paper is organized as follows. Section 2 suggests an improved problem formulation. The solution procedure is illustrated in Section 3. Concluding remarks are provided in Section 4. Points of disagreement are marked in italics.
## 2 Improved Problem Formulation
We adopt the model of distributed computing as proposed in . In this model host computers are treated as node. Each node $`i\{1,2,\mathrm{}n\}`$ is capable of executing any incoming job k. Arrival rate of job at node i is $`\varphi _i`$, and total arrival rate of jobs on the system is $`\mathrm{\Phi }=_{i=1}^n\varphi _i`$. When a job k arrive at a node i, it may be executed at node i itself or transferred to another node j, depending on the distribution of loads at that time. If job is transferred to node j, it is executed at node j without being transferred to any other host, and results of execution is transferred back to node i so that the user submitting the job gets the illusion of the job being executed at the host i. Processing rate of job at node i is $`\beta _i`$ and transfer rate of jobs from node i to node j is $`x_{ij}`$. Therefore, $`\lambda =_{i=1}^n_{j=1}^nx_{ij}`$ is the total traffic on the network.
It is also assumed that hosts may be heterogeneous (different CPU speed, memory and swap availability etc.) and speed of execution depends on the instantaneous load of the host. Therefore the instantaneous marginal delay due to local processing of a job k at host i depends on host i as well as load on the host i, and is denoted by $`F_i(\beta i)`$. So the mean response time of all the jobs on the network due to local processing is given by $`_{i=1}^n\frac{\beta _i}{_{i=1}^n\beta _i}F_i(\beta i)=_{i=1}^n\frac{\beta _i}{\beta }F_i(\beta i)`$.
In the distributed set up, another major component of mean response time is the network delay. The incremental communication delay for transfer of a packet can be modeled as $`G_{ij}(๐ฑ)`$, where $`๐ฑ`$ is the vector of load transfer. Note that we are assuming that the network delay due to sending the response back is negligible. In case it is not so, the argument of G should be suitably replaced. <sup>1</sup><sup>1</sup>1If we assume that $`y_{ij}`$ is the amount of flow of results from node i to node j, then a more accurate model would have network delay given by: $`_{i=1}^n_{j=1}^n\frac{x_{ij}+y_{ji}}{_{i=1}^n_{j=1}^n(x_{ij}+y_{ji})}G_{ij}(๐ฑ+๐ฒ)`$, where $`๐ฒ`$ is the vector of additional traffic due to response packets. So the contribution of network delay to mean response time is given by $`_{i=1}^n_{j=1}^n\frac{x_{ij}}{_{i=1}^n_{j=1}^nx_{ij}}G_{ij}(๐ฑ)`$.
\[It is to be noted that in as well as in , denominator used was $`\mathrm{\Phi }`$ which is not equal to $`_{i=1}^n_{j=1}^nx_{ij}`$ . Hence, their formulations did not use average communication delay, contrary to their claims. Also the claim in that $`๐ฑ`$ in $`G(๐ฑ)`$ can include response packets is wrong as in that case total flow no longer remains $`\lambda `$.\]
Hence the problem of load balancing can be formulated as:
Minimize
$$\underset{i=1}{\overset{n}{}}\frac{\beta _i}{_{i=1}^n\beta _i}F_i(\beta i)+\underset{i=1}{\overset{n}{}}\underset{j=1}{\overset{n}{}}\frac{x_{ij}}{_{i=1}^n_{j=1}^nx_{ij}}G_{ij}(๐ฑ)$$
subject to
$$\underset{i=1}{\overset{n}{}}\beta _i=\underset{i=1}{\overset{n}{}}\varphi _i,$$
$$\beta _i+\underset{j=1}{\overset{n}{}}x_{ij}=\varphi _i+\underset{j=1}{\overset{n}{}}x_{ji},i=1,2,\mathrm{},n$$
$$\lambda =\underset{i=1}{\overset{n}{}}\underset{j=1}{\overset{n}{}}x_{ij},$$
$$\beta _i0,i=1,2,\mathrm{},n,$$
$$x_{ij}0,i,j=1,2,\mathrm{},n,$$
$$x_{ii}=0,i=1,2,\mathrm{},n,$$
(1)
Note that first two constraints ensure that all incoming load is processed, and flow balance holds true. As in , we assume that $`F_i(\beta _i)`$ are increasing convex functions and $`G_{ij}(๐ฑ)`$ is a non-decreasing convex function. Ethernet and satellite are quite common network infrastructure. Due to the type of data transmission means on these, the network delay for transfer of a packet from node i to node j depends on the total network load. Henceforth, we also assume that $`G_{ij}(๐ฑ)`$ depends on the total traffic $`\lambda `$ on the network but not on any component of it. So the second term in the objective function can be written as $`_{i=1}^n_{j=1}^n\frac{x_{ij}}{\lambda }G_{ij}(\lambda )`$.
## 3 Solution of the Problem
Solution to the above problem can be expressed in terms of disjoint set of nodes. We classify the nodes as source (idle or active), neutral and sinks as in . A node $`i`$ is said to be source if $`x_{ji}=0j`$ and $`jx_{ij}>0`$. A source $`i`$ is idle if $`\beta _i=0`$, else it is active. A node $`i`$ is said to be sink if $`x_{ij}=0j`$ and $`jx_{ji}>0`$. For a neutral node $`i`$, $`x_{ij}=0j`$ and $`x_{ji}=0j`$. We denote the set of sinks, idle sources, active sources and neutrals by $`S,R_d,R_a`$ and $`N`$ respectively.
Tantawi and Towsley claimed that triangular inequality of network delays (i.e., $`G_{ij}(\lambda )G_{ik}(\lambda )+G_{kj}(\lambda )`$) was sufficient condition for solution of the problem as formulated by them to yield only source, sink and neutral nodes. We illustrate in the following theorem that the claim is wrong.
###### Theorem 1
The triangular inequality of network delay and non-decreasing property of $`\frac{G(\lambda )}{\lambda }`$ are sufficient conditions for Problem 1 to yield nodes which are either sources, sinks or neutrals.
Proof: For a given incoming load $`\beta `$ is constant, and hence we just need to show that the stated conditions are sufficient for a flow rate assignment $`๐ฑ`$ with nodes either sinks, sources or neutral nodes to minimize the communication term $`C=_{i=1}^n_{j=1}^n\frac{x_{ij}}{\lambda }G_{ij}(\lambda )`$. Let us suppose that $`i(๐ฑ)`$ denotes the number of nodes that are neither sinks, sources nor neutrals for the flow rate assignment $`๐ฑ`$. Therefore we need to show that $`i(๐ฑ)=0`$. Of all the flow rate assignment minimizing communication delay, choose the flow rate assignment $`๐ฑ`$ which has maximum $`i(๐ฑ)`$ and it is strictly positive. \[In Proof of Theorem 1 in , $`๐ฑ`$ was chosen to have minimum $`i(๐ฑ)`$ from candidate solutions minimizing communication delay. And it was proved that $`i(๐ฑ)=0`$. But this does not rule out the possibility that there can be other optimal solutions with $`i(๐ฑ)>0`$. So it only proves that there is at least one flow rate assignment minimizing the objective function for which nodes are either sources, sinks or neutrals. \] Since $`i(๐ฑ)>0`$, there is at least one node $`k`$ and a pair $`l`$ and $`m`$ such that $`x_{lk}>0`$ and $`x_{km}>0`$. Consider a new assignment $`๐ฑ^{}`$ which differs from $`๐ฑ`$ only in the following.
$$x_{lk}^{}=x_{lk}\mathrm{min}\{x_{lk},x_{km}\},$$
$$x_{km}^{}=x_{km}\mathrm{min}\{x_{lk},x_{km}\},$$
$$x_{lm}^{}=x_{lm}+\mathrm{min}\{x_{lk},x_{km}\}.$$
Let the new Communication delay be $`C^{}`$. So,
$$C^{}=\underset{i=1}{\overset{n}{}}\underset{j=1}{\overset{n}{}}\frac{x_{ij}^{}}{\lambda ^{}}G_{ij}(\lambda ^{})=$$
$$x_{lk}^{}\frac{G_{lk}(\lambda ^{})}{\lambda ^{}}+x_{km}^{}\frac{G_{km}(\lambda ^{})}{\lambda ^{}}+x_{lm}^{}\frac{G_{lm}(\lambda ^{})}{\lambda ^{}}+(\underset{i=1}{\overset{n}{}}\underset{j=1}{\overset{n}{}}x_{ij}\frac{G_{ij}(\lambda ^{})}{\lambda ^{}}(i,j)(l,k),(k,m),(l,m)\}).$$
Since $`\lambda ^{}\lambda `$ and $`\frac{G(\lambda )}{\lambda }`$ is non-decreasing,
$$C^{}C\{x_{lk}^{}\frac{G_{lk}(\lambda ^{})}{\lambda ^{}}x_{lk}\frac{G_{lk}(\lambda )}{\lambda }\}+\{x_{km}^{}\frac{G_{km}(\lambda ^{})}{\lambda ^{}}x_{km}\frac{G_{km}(\lambda )}{\lambda }\}$$
$$+\{x_{lm}^{}\frac{G_{lm}(\lambda ^{})}{\lambda ^{}}x_{lm}\frac{G_{lm}(\lambda )}{\lambda }\}(x_{lk}^{}x_{lk})\frac{G_{lk}(\lambda )}{\lambda }$$
$$+(x_{km}^{}x_{km})\frac{G_{km}(\lambda )}{\lambda }+(x_{lm}^{}x_{lm})\frac{G_{lm}(\lambda )}{\lambda }$$
$$=\frac{1}{\lambda }[\mathrm{min}\{x_{lk},x_{km}\}(G_{lk}(\lambda )+G_{km}(\lambda )G_{lm}(\lambda ))]0.$$
Therefore $`i(๐ฑ^{})<i(๐ฑ)`$, a contradiction.
\[in Proof of Theorem 1 in , with new $`๐ฑ^{}`$ total network load no longer remains $`\lambda `$. Also it is possible to have $`i(๐ฑ^{})=i(๐ฑ)2`$, and not always $`i(๐ฑ^{})=i(๐ฑ)1)`$.\] โ
As for broadcast network, delay can be assumed to be independent of source-destination pair, $`G_{ij}(\lambda )`$ is independent of i and j. Moreover, since $`_{i=1}^n\beta _i=_{i=1}^n\varphi _i`$ and total incoming load is constant, the objective function in Problem 1 is equivalent to
$$\underset{i=1}{\overset{n}{}}\beta _iF_i(\beta _i)+\underset{i=1}{\overset{n}{}}\beta _iG(\lambda ).$$
For simplifying the solution procedure, let $`u_i=_{j=1}^nx_{ji},v_i=_{j=1}^nx_{ij}`$. So Problem 1 can be written as:
Minimize
$$\underset{i=1}{\overset{n}{}}\beta _iF_i(\beta _i)+\underset{i=1}{\overset{n}{}}\beta _iG(\underset{i=1}{\overset{n}{}}v_i)$$
subject to
$$\beta _i+v_i=\varphi _i+u_i,i=1,2,\mathrm{},n,$$
$$\underset{i=1}{\overset{n}{}}v_i\underset{i=1}{\overset{n}{}}u_i=0,i=1,2,\mathrm{},n$$
$$\beta _i0,i=1,2,\mathrm{},n,$$
$$u_i0,i=1,2,\mathrm{},n,$$
$$v_i0,i=1,2,\mathrm{},n.$$
(2)
Note that $`_{i=1}^n\beta _i=_{i=1}^n\varphi _i=\mathrm{\Phi }`$ is constant. After eliminating the first constraint, problem 2 has the following form.
Minimize
$$\underset{i=1}{\overset{n}{}}(u_iv_i+\varphi _i)F_i(u_iv_i+\varphi _i)+\mathrm{\Phi }G(\underset{i=1}{\overset{n}{}}v_i)$$
subject to
$$\underset{i=1}{\overset{n}{}}v_i\underset{i=1}{\overset{n}{}}u_i=0,i=1,2,\mathrm{},n$$
$$u_iv_i+\varphi _i0,i=1,2,\mathrm{},n,$$
$$u_i0,i=1,2,\mathrm{},n,$$
$$v_i0,i=1,2,\mathrm{},n.$$
(3)
###### Theorem 2
The optimal solution to the above problem can be expressed as:
$$f_i(\beta _i)\alpha +\beta G^{}(\lambda ),\beta _i=0(iR_d),$$
$$f_i(\beta _i)=\alpha +\beta G^{}(\lambda ),\mathrm{\hspace{0.33em}0}<\beta _i<\varphi _i(iR_a),$$
$$\alpha f_i(\beta _i)\alpha +\beta G^{}(\lambda ),\beta _i=\varphi _i(iN),$$
$$f_i(\beta _i)=\alpha ,\beta _i>\varphi _i(iS),$$
subject to total flow constraint,
$$\underset{iR_a}{}f_i^1(\alpha +\mathrm{\Phi }G^{}(\lambda ))+\underset{iN}{}\varphi _i+\underset{iR_a}{}f_i^1(\alpha )=\mathrm{\Phi },$$
where
$$\lambda =\underset{iS}{}(f_i^1(\alpha )\varphi _i).$$
Proof: The Lagrangean function for this problem is
$$L(๐ฎ,๐ฏ,\alpha ,\gamma ,\psi ,\eta )=\underset{i=1}{\overset{n}{}}(u_iv_i+\varphi _i)F_i(u_iv_i+\varphi _i)+\mathrm{\Phi }G(\underset{i=1}{\overset{n}{}}v_i)$$
$$+\alpha (\underset{i=1}{\overset{n}{}}v_i\underset{i=1}{\overset{n}{}}u_i)+\underset{i=1}{\overset{n}{}}\gamma _i(u_iv_i+\varphi _i)+\underset{i=1}{\overset{n}{}}\psi _iu_i+\underset{i=1}{\overset{n}{}}\eta _iv_i.$$
It can be shown easily that objective function and the constraints satisfy conditions for applying Karush-Kuhn-Tucker (KKT) conditions. Let $`f_i(\beta _i)=F_i(\beta _i)+\beta _i\frac{F_i(\beta _i)}{\beta _i}`$. Applying KKT conditions, we get the following set of equations.
$$\frac{\delta L}{\delta u_i}=f_i(u_iv_i+\varphi _i)\alpha +\gamma _i+\psi _i=0,i=1,2,\mathrm{},n,$$
(4)
$$\frac{\delta L}{\delta v_i}=f_i(u_iv_i+\varphi _i)+\mathrm{\Phi }G^{}(\underset{i=1}{\overset{n}{}}v_i)+\alpha \gamma _i+\eta _i=0,i=1,2,\mathrm{},n,$$
(5)
$$\frac{\delta L}{\delta \alpha }=\underset{i=1}{\overset{n}{}}u_i+\underset{i=1}{\overset{n}{}}v_i=0,$$
(6)
$$u_iv_i+\varphi _i0,\gamma _i(u_iv_i+\varphi _i)=0,\gamma _i0,i=1,2,\mathrm{},n,$$
(7)
$$u_i0,\psi _iu_i=0,\psi _i0,i=1,2,\mathrm{},n,$$
(8)
$$v_i0,\eta _iv_i=0,\eta _i0,i=1,2,\mathrm{},n.$$
(9)
Let us consider two cases separately, $`u_iv_i+\varphi _i=0`$ and $`u_iv_i+\varphi _i>0`$. Case 1: $`u_iv_i+\varphi _i=0`$, i.e., $`\beta _i=0`$.
Again we consider two sub cases:
1A: $`\varphi _i>0`$.
In this case $`v_i>0`$, and consequently from Equation 9 $`\eta _i=0`$.So from Equation 5
$$f_i(\beta _i)\alpha +\beta G^{}(\lambda ).$$
Also for this case, $`\beta _i=0,\varphi _i>0`$. It can easily be noticed that these are idle source nodes.
1B: $`\varphi _i=0`$.
It is obvious that in this case $`u_i=v_i`$. From Equation 4 and Equation 5,
$$\mathrm{\Phi }G^{}(\lambda )+\gamma _i+\eta _i=0.$$
But by assumption $`\mathrm{\Phi }G^{}(\lambda )>0`$ and from Equation 7 and Equation 9 $`\gamma _i0,\eta _i0`$. This implies atleast one of $`\gamma _i`$ and $`\eta _i`$ is strictly negative, and correspondingly from Equation 7 or Equation 9 either $`u_i`$ or $`v_i`$ is zero. Therefore $`u_i=v_i=0`$. Hence, from Equation 5 $`f_i(\beta _i)\alpha .`$ Also $`\beta _i=0,\varphi _i=0`$.
These nodes correspond to neutral nodes without external load.
Case 2: $`u_iv_i+\varphi _i>0`$.
From Equation 7, $`\gamma _i=0`$. Now consider three subcases:
2A: $`v_i>0,u_i=0`$.
Since $`v_i>0`$, from Equation 9 $`\eta _i=0`$. Hence from Equation 5, we get
$$f_i(\beta _i)=\alpha +\beta G^{}(\lambda ).$$
(10)
For this case, $`0<\beta _i<\varphi _i`$. Therefore these nodes correspond to active sources.
2B: $`v_i=0,u_i>0`$.
Since $`u_i>0`$, from Equation 8 $`\psi _i=0`$. Hence from Equation 4,
$$f_i(\beta _i)=\alpha .$$
In this case $`\beta _i>\varphi _i`$. These are sink nodes.
2C. $`u_i=v_i=0`$.
Since $`\psi _i0`$, from Equation 4 we get $`f_i(\beta _i)\alpha `$. Similarly, since $`\eta _i0`$, from Equation 5 we get $`f_i(\beta _i)\alpha +\beta G^{}(\lambda )`$. Here $`\beta _i=\varphi _i`$, and hence nodes are neutrals.
To solve the system completely, we need to find out value of $`\alpha `$. From Equation 6, $`_{i=1}^n\beta _i=\mathrm{\Phi }`$. Substituting values of $`\beta _i`$ from the solution nodes in the expression we get $`_{iR_a}f_i^1(\alpha +\mathrm{\Phi }G^{}(\lambda ))+_{iN}\varphi _i+_{iR_a}f_i^1(\alpha )=\mathrm{\Phi }`$. We still need value of $`\lambda `$ to solve the above equation to find out $`\alpha `$ for the incumbent solution. Fortunately $`\lambda =_{i=1}^nu_i=_{iS}u_i=_{iS}(\beta _i+v_i\varphi _i)=_{iS}(\beta _i\varphi _i)=_{iS}(f_i^1(\alpha )\varphi _i)`$. Now we are at a position to find out $`\alpha `$ from the above equation.
\[There are some mistakes in the proof of corresponding theorem in which we presume to be typographical error.\] โ
The result of the above theorem can be intuitively explained as illustrated below. All the sink nodes have the lowest and equal incremental node delay($`\alpha `$). Loads are shared among the sink nodes such that all have equal incremental node delay. For all the active source nodes the incremental node delay equals sum of incremental node delay of sink nodes and total incremental communication delay at the present load ($`\beta `$). If the incremental node delay at a source is greater than this value then it is profitable, from the perspective of the objective to reduce mean response time, to transfer all the load of the source to the sink nodes, and these nodes correspond to idle source nodes. Note that load distribution among the active source nodes are such that all of them have equal incremental node delay($`\alpha +\beta G^{}(\lambda )`$), and these nodes process portion of their incoming load ($`\varphi _i`$) transferring the remaining loads to sinks. However, if at a node the incremental node delay is less than the above value, then it is not profitable to send any portion of the load to sink node and thereby incur communication delay. All the incoming load at such a node is processed by itself, and it is a neutral node. The above solution is soothing to practitionersโ eyes also. The optimal solution suggests a distribution such that nodes with lower incremental node delay are utilized more extensively than others.
## 4 Conclusion
This article points out a number of inconsistencies in , and improves on them. One of the nice aspects of the results in the present article is that the parametric analysis as well as single-point load balancing algorithm as suggested in can be applied with suitable modification of expressions from the solution given in the previous theorem. It is also perceivable that due to the simplicity of the expression, the present procedures can not be more difficult than those in .
Fortunately, the present finding does not affect the results of . As mentioned before, Ross et al. had considered a general problem consisting of generic as well as dedicated jobs, and included the task of scheduling jobs at each host. The authors showed that problem was separable over scheduling decisions. They also showed that given an allocation of the jobs on the hosts, the task of scheduling can be solved as a polymatroid optimization problem. The present results in this paper only changes the allocation of the jobs on the hosts and does not violate any of the assumptions made in .
As noted in , dynamic load balancing can yield better response time. Most of the dynamic load balancing tool uses a decision making unit (called load balancer) which periodically monitors load on hosts. Also hosts can decide to inform of any exigency when load on them exceed a high level threshold or goes below a low level threshold (both these thresholds can be dynamic). Load balancer then decides on new load distribution and executes the transfer. The present result can be used for dynamic load balancing. The expected values of $`\alpha `$ and $`\alpha +\beta G^{}(\lambda )`$ can serve as lower and higher thresholds. The load balancer can keep track of incremental node delay (a function of allocated load on the node. See for some illustrative measures of load) of hosts as well as incremental communication delay. When it receives any receiver-initiated request for a load, generated by load on the host going below the lower threshold), it can invoke single-point algorithm to reallocate the load. Similarly, it can redistribute load when it receives any sender-initiated request, generated by load on the host going above the upper threshold.
|
warning/0006/hep-th0006090.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
It is now well understood that the low energy dynamics of a single $`Dp`$-brane is described by the Dirac-Born-Infeld action, combined with a Wess-Zumino term which describes the coupling of the brane to the background RR fields and gravity curvature. This action is consistent with T-duality, in the sense that it can be obtained from the ten dimensional BI action by dimensional reduction. It admits an extension to the $`\kappa `$โinvariant action for both flat and general type II supergravity backgrounds. This action is effective since it describes only the dynamics of the massless excitations of the brane. It is low energy in the sense that it is valid only for slowly varying gauge fields, i.e. there are corrections to this action of order $`\mathrm{O}(l_s^3F)`$. However, the abelian DBI action incorporates corrections to all orders in the field strength. The physical degrees of freedom of the DBI action are manifest in the static gauge and they correspond to the degrees of freedom of a $`p+1`$-dimensional $`U(1)`$ gauge theory, $`9p`$ transverse scalar fields and their fermionic partners.
On the other hand, the low energy dynamics of a system of $`Dp`$ branes is still not well understood. The lowest energy description for a system of $`Dp`$-branes is given by the maximally supersymmetric $`U(N)`$ Yang-Mills theory in $`p+1`$ dimensions. But the full effective action is unknown. One of the main problems is the following. In the abelian case, one can uniquely split the abelian action into those parts which do and those parts which do not depend on the derivative of $`F`$. For the system of branes that is not the case since $`[D_a,D_b]F_{cd}=[F_{ab},F_{cd}]`$. Hence, one is allowed to โtradeโ $`DF`$ terms for commutators of $`F`$ terms. To overcome this ambiguity, when talking about the nonabelian Born-Infeld action (hereafter referred to as NBI action) one means only the part of the effective action which does not contain any commutators of $`F`$. Tseytlin has defined the NBI action as that part of the full open string effective action which is completely symmetric in all factors of $`F`$ in each monomial $`\mathrm{tr}(\mathrm{F}\mathrm{}\mathrm{F})`$. The full open string effective action of the $`Dp`$-brane is then given by the sum of the following terms: the NDBI action, the $`F^n`$ terms containing factors of commutators of $`F`$, and the terms with symmetrised covariant derivatives. Hence a discrepancy between the predictions of the full string theory and the predictions of the nonabelian DBI action (such as found in ) is not unexpected.<sup>2</sup><sup>2</sup>2Supersymmetrisation of this action is also a nontrivial problem .
We will not consider here questions about the validity of the NDBI action and proposals for possible correction terms which should be added to this action. Instead, we will focus on rules for efficient calculation with the existing NDBI action. As given, Tseytlinโs proposal for the nonabelian action allows one to work only in a power expansion of the field strength $`F`$. However, one would wish to introduce a closed-form notation, which would allow one to do all calculations without writing explicit power expansions. Of course, the new objects that we define are such that at each stage of the calculation one can rewrite all results in a power expansion. In the literature many calculations have been done in closed form, implicitly introducing Lie algebra valued functions. However, strict definitions as well as precise rules for how one should do calculations with these functions have not been given. The first part of this paper aims to fill this gap. We also point out the limitations of this approach. Namely, after certain operations one loses the ability to further express things in a closed form.
In the second part of the paper we apply these rules to find a particular solution of the NDBI action. It was recently found that five dimensional $`N=2`$ <sup>3</sup><sup>3</sup>3We use language of the five dimensional spinors, i.e. $`N=2`$ theory has 16 real supercharges. supersymmetric Yang-Mills-Higgs theory with gauge group $`SU(N)`$ broken down to the maximal torus $`U(1)^{(N1)}`$ admits a new type of nonsingular instanton solution. This instanton, despite the broken gauge group and broken scale invariance of the theory is not singular due to the fact that it carries electric charge under the unbroken Abelian subgroup. This charge supports the instanton against collapse. Since the instanton carries both topological charge $`k`$ (winding number) and electric charge $`q`$, the authors of have christened it a dyonic instanton. It was shown in that this configuration is supersymmetric, preserving 1/4 of the supersymmetries. We will show that this solution solves the full non-abelian DBI action for the $`D4`$ brane. We interpret this solution, for the case of a $`SU(2)`$ gauge group, as the superposition of a $`D0`$ brane with a fundamental string connecting two parallel $`D4`$ branes. The geometry of this configuration is interesting and new compared to the shape of the brane with monopole solutions given in . Due to the presence of the delocalised source of the electric field on the brane, a new effect of blowing-up of the brane appears. We also comment on the problem that one encounters when trying to abelianise the solution.
## 2 Systematics of the NDBI calculus
Before discussing the nonabelian dyonic instantons, we will now first present a systematic method for closed-form calculations involving the NDBI action. As proposed by Tseytlin the NBI action for the $`D9`$ brane is given by
$`L_{NBI}(F)=Str\sqrt{det(\eta _{\mu \nu }+T^1F_{\mu \nu })}.`$ (2.1)
where $`\mu ,\nu =0,\mathrm{},9`$ and $`T=\frac{1}{2\pi \alpha ^{}}`$. The symbol Str in (2.1) is defined in the following way. By expanding the abelian BI action as
$`\sqrt{det(\eta _{\mu \nu }+F_{\mu \nu })}={\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}C^{\mu _1\nu _1\mathrm{}\mu _{2k}\nu _{2k}}F_{\mu _1\nu _1}\mathrm{}F_{\mu _{2k}\nu _{2k}}.`$ (2.2)
one defines the Lorentz tensors $`C^{\mu _1\nu _1\mathrm{}\mu _{2k}\nu _{2k}}`$. Then the NBI action is defined as a power series
$`Str\sqrt{det(\eta _{\mu \nu }I+F_{\mu \nu })}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}d_{a_1\mathrm{}a_{2k}}C^{\mu _1\nu _1\mathrm{}\mu _{2k}\nu _{2k}}F_{\mu _1\nu _1}^{a_1}\mathrm{}F_{\mu _{2k}\nu _{2k}}^{a_{2k}}.`$ (2.3)
Here the tensors
$`D_{a_1\mathrm{}a_p}`$ $`=`$ $`{\displaystyle \frac{1}{n!}}(T_{a_1}\mathrm{}T_{a_p}+allpermutations)`$
$`d_{a_1\mathrm{}a_p}`$ $`=`$ $`tr(D_{a_1\mathrm{}a_p})`$ (2.4)
are totally symmetric, adjoint action invariant tensors of the gauge group.<sup>4</sup><sup>4</sup>4A major achievement of this action is that it reproduces correctly all terms, up to the $`\alpha ^2F^4`$, from string theory calculations. Strictly speaking, given this definition of Str, one can only work with a power expansion form of the action. However, extending the Tseytlin definition as follows allows one to do calculations under the Str without writing out a power expansion. Following definition (2.3), one first introduces a new type of object: a function of the the Lie algebra variables, $`\widehat{f}`$.
Let $`f=f(A_1,\mathrm{},A_n)`$ be a $`C^{\mathrm{}}`$ function with the expansion
$`f={\displaystyle \underset{\genfrac{}{}{0pt}{}{k=0}{h_1^k+\mathrm{}+h_n^k=k}}{\overset{\mathrm{}}{}}}f^{h_1^k\mathrm{}.h_n^k}A_1^{h_1^k}\mathrm{}A_n^{h_n^k}.`$ (2.5)
Then the corresponding function $`\widehat{f}`$ over Lie algebra variables $`\widehat{A}_i=\widehat{A}_i^aT^a`$, $`i=1,\mathrm{},n`$ is defined as
$`\widehat{f}{\displaystyle \underset{\genfrac{}{}{0pt}{}{k=0}{h_1^k+\mathrm{}+h_n^k=k}}{\overset{\mathrm{}}{}}}f^{h_1^k\mathrm{}.h_n^k}\widehat{A}_1^{๐_\mathrm{๐}}\mathrm{}\widehat{A}_n^{๐_๐ค}D_{๐_\mathrm{๐}\mathrm{}๐_๐ค}`$ (2.6)
where a $`D`$-tensor is defined as in (2).
Next, we define operations with hat functions. Additivity and multiplication by a scalar is naturally induced from the algebra of the โabelianโ functions $`C^{\mathrm{}}`$. Note that the induced addition for the hat functions is commutative: $`\widehat{f}+\widehat{g}=\widehat{g}+\widehat{f}`$. Multiplication in $`C^{\mathrm{}}`$ induces a natural multiplication among the hat functions as follows. For simplicity we will talk below about functions of a single Lie algebra variable, and comment on the generalisation to the case of several variables only where it has specific features. For functions $`\widehat{g}(A)`$ and $`\widehat{l}(A)`$ the first guess for the induced product $`\widehat{g}(A)\widehat{l}(A)`$ might be
$`\widehat{g}(\widehat{A})\widehat{l}(\widehat{A}){\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}g^kl^m\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\widehat{A}^{๐_๐ค}\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\widehat{A}^{๐_๐ฆ}D_{๐_\mathrm{๐}\mathrm{}๐_๐ค}D_{๐_\mathrm{๐}\mathrm{}๐_๐ฆ}`$ (2.7)
However, this is not a well defined operation since the product of two hat functions is not a hat function. If the abelian counterpart functions satisfy $`f=lg`$, then this product does not imply $`\widehat{f}=\widehat{l}\widehat{g}`$ since
$`D_{๐_\mathrm{๐}\mathrm{}๐_๐ค}D_{๐_\mathrm{๐}\mathrm{}๐_๐ฆ}D_{(๐_\mathrm{๐}\mathrm{}๐_๐ค}D_{๐_1\mathrm{}๐_m)}`$ (2.8)
Only in the case of a single variable $`A`$ does the equality $`\widehat{f}=\widehat{l}\widehat{g}`$ hold, since $`\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\widehat{A}^{๐_๐ค}\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\widehat{A}^{๐_๐ฆ}`$ is a totally symmetric expression, and hence induces symmetrisation in the product of $`D`$-tensors. The approach adopted in was to treat all hat functions under the Str as abelian. This means that the product (2.7) is modified to
$`\widehat{g}(\widehat{A})\widehat{l}(\widehat{A}){\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}g^kl^m\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\widehat{A}^{๐_๐ค}\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\widehat{A}^{๐_๐ฆ}D_{(๐_\mathrm{๐}\mathrm{}๐_๐ค}D_{๐_\mathrm{๐}\mathrm{}๐_๐ฆ)}`$ (2.9)
where we have introduced additional symmetrisation over products of $`D`$-tensors. This product is obviously closed and moreover respects associativity and is commutative. When supplied with operations for scalar multiplication, addition and multiplication, the set of hat functions forms an algebra.
Another property that we will use is that variations of the hat functions can be taken as in the abelian case, i.e. one can take variations without worrying about the gauge group structure (as long as we are assuming the โ * โ product)
$`\delta \widehat{f}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}kg^k\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\delta \widehat{A}^{๐_๐ค}D_{๐_\mathrm{๐}\mathrm{}๐_๐ค}`$ (2.10)
$`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}kg^k\widehat{A}^{๐_\mathrm{๐}}\mathrm{}\widehat{A}^{๐_{๐ค\mathrm{๐}}}D_{๐_\mathrm{๐}\mathrm{}๐_{๐ค\mathrm{๐}}}\delta \widehat{A}`$ (2.11)
$``$ $`{\displaystyle \frac{\delta \widehat{f}}{\delta \widehat{A}}}\delta \widehat{A}`$ (2.12)
Hence the variation of the action (2.1) is
$$\delta L=Str[\sqrt{det\widehat{H_{\xi \rho }}}(\widehat{H}^1)^{\mu \nu }\delta F_{\mu \nu }],$$
(2.13)
where $`\widehat{H}_{\mu \nu }\eta _{\mu \nu }+F_{\mu \nu }`$ and $`\delta F_{\mu \nu }=D_\mu \delta A_\nu D_\nu \delta A_\mu `$ . Since for $`F`$ (and quantities derived from it by T-duality) there is no distinction between ordinary and hat functions, we will henceforth write them without the hat.
All we have done up to now referred only to the $`D9`$ brane. To get the NDBI action for any $`Dp`$-brane, one goes to the static gauge, identifies world volume coordinates (along which we want to T-dualise) with background coordinates and applies T-duality 9-$`p`$ times to (2.1). After T-duality we get the action
$`S_p`$ $`=`$ $`T_p{\displaystyle d^{p+1}xStr\sqrt{det}(\eta _{\mu \nu }+T^1F_{\mu \nu })}`$ (2.14)
$`=`$ $`T_p{\displaystyle d^{p+1}xStr\left[\sqrt{det\widehat{๐ข}_{mn}}\sqrt{det\widehat{๐ข}_{rs}}\right]},`$
$`\widehat{๐ข}_{mn}\eta _{mn}+\widehat{๐ข}^{rs}(X)D_mX_rD_nX_s+T^1F_{mn},\widehat{๐ข}_{rs}(X)\delta _{rs}iT[X_r,X_s],`$ (2.15)
where indices $`m,n=0,\mathrm{},p`$ and $`r,s,=p+1,\mathrm{},9`$. It is important to note that all hat functions under the Str are functions of the variables $`F`$ and quantities derived from it by T-duality,
$`F_{rs}`$ $`=`$ $`iT^2[X_r,X_s],`$
$`F_{mr}`$ $`=`$ $`TD_mX_r=T(_mX_ri[A_m,X_r]),`$
$`F_{mn}(A)`$ $`=`$ $`_mA_n_nA_mi[A_m,A_n].`$ (2.16)
One is $`\mathrm{๐ง๐จ๐ญ}`$ associating Lie algebra generators to $`X`$โs and $`A`$โs, but just to $`F`$โs under Str. Hat functions are by definition functions of Lie algebra variables $`F`$, not $`A`$ or $`X`$. As soon as we pass from $`F`$โs to $`A`$โs or $`X`$โs we lose the ability to write expansions in closed hat function form. In order to go from the first to the second equality in (2.14) we have used
$`det\left(\begin{array}{cc}A& B\\ C& D\end{array}\right)`$ $`=`$ $`det(D)det(ABD^1C)`$ (2.19)
where here it is understood that $`det`$ is the hat determinant. If we compare the action (2.14) with the abelian case of the appropriate dimension we see that the nonabelian action for $`Dp`$-branes is not just the abelian action with the replacement $`D`$ and Str introduced. There are additional commutation terms $`[X,X]`$ and $`[A,X]`$ in the nonabelian action.
Another operation that one often performs under $`Str`$ is
$`tr(\widehat{f}\delta F_{mn})=tr\left(\widehat{f}(D_m\delta A_nD_n\delta A_m)\right).`$ (2.20)
Here in the second equation we have passed from the โ * โ product to the ordinary product. We can do this due to the fact that $`F`$ carries only one Lie algebra index. In each term of the Str we can always cyclically permute the $`T^a`$ generator (corresponding to the $`F`$) under the trace to the right. So this operation of passing from โ * โ to the ordinary product is possible only under the Str and only for the expressions of the form $`tr(fl)`$ where $`f=f^aT^a`$ and $`l`$ arbitrary.
Next we want to define the covariant derivative of the hat function. In order to do this we first note that any hat function $`\widehat{f}`$, valued in the $`SU(N)`$ gauge group, can be written as
$`\widehat{f}=f^aT^a+f^0I`$ (2.21)
where $`I`$ is the identity matrix. This follows from the property of the $`SU(N)`$ generators $`T^aT^b=\frac{c}{2}\delta ^{ab}+\frac{1}{2}d_c^{ab}T^c+\frac{i}{2}f_c^{ab}T^c`$. Hence, any product of generators can be expressed as linear combination of the identity matrix and the generators. The covariant derivative is then defined as
$`D_m\widehat{f}_mf^aT^a+_mf^0+f^a[A_m,T^a]`$ (2.22)
Although it is not obvious from the definition (2.21) that the covariant derivative of a hat function is itself a hat function, one can prove this easily by writing $`\widehat{f}`$ explicitly in expanded form. However, the abelian counterpart function of $`D\widehat{f}`$ is not obvious. Generically, $`D\widehat{f}`$ is of the form: $`D\widehat{f}=\widehat{f}+\widehat{H}`$ where $`\widehat{H}`$ is not directly expressible in terms of $`\widehat{f}`$. This is in contrast to ordinary nonabelian gauge theory where the analogue of $`\widehat{H}`$ is a commutator of a gauge potential and a function (here $`\widehat{H}A\widehat{f}\widehat{f}A`$). In order to calculate $`\widehat{H}`$ explicitly one has to use the full expanded form. As long as we treat $`D\widehat{f}`$ as a single object (not worrying about its explicit form) we can treat all expressions as abelian ($`D\widehat{f}`$ is a hat function). However, if we want to perform some calculation which depends explicitly on the structure of $`D\widehat{f}`$ (for example, to calculate $`\delta D\widehat{f}`$) we must work with the definition of $`D\widehat{f}`$ in expanded form.
Now we can do a partial integration of (2.20) so that it becomes
$`tr(\widehat{f}\delta F_{mn})=tr((D_m\widehat{f})\delta A_n(D_n\widehat{f})\delta A_m).`$ (2.23)
In summary, as long as we are dealing with hat functions under Str as functions of the Lie algebra variable $`F`$ and perform operations which are well defined for them, i.e. operations which send hat functions into hat functions (multiplication by a scalar, addition, multiplication and partial differentiation) we can treat all expressions under the symmetrised trace as abelian. On the other hand, the change of variables from $`F`$ to $`A`$ and explicit calculation with covariant derivative of hat function is possible only in the expansion form.
## 3 Dyonic instanton as solution of the nonabelian DBI action
In a recent paper by Lambert and Tong , it was shown that Yang-Mills-Higgs theory broken to the $`U(1)`$ subgroup admits a new type of the instanton solution, the dyonic instanton. This instanton is a BPS instanton embedded in the 4+1, $`SU(2)`$ Yang-Mills-Higgs theory broken to $`U(1)`$. This instanton is charged with respect to the unbroken $`U(1)`$ gauge group. We will now apply the rules for the hat functions in order to show that the dyonic instanton is a solution to the NDBI action.
Since the dyonic instanton is characterised by only one transverse scalar being nonzero, the action (2.14) in this case simplifies greatlyโall commutators vanish identically. In what follows all functions are understood to be hat functions and all products โ * โ products, but to simplify notation we will not write them explicitly.
In order to show that this configuration satisfies the equations of motion of the NDBI action it is convenient to first rewrite the action (2.14) as the non-abelian BI action. Using the T-duality rules (2), we can rewrite the NDBI action for the D4 brane as
$`L={\displaystyle \frac{1}{(2\pi \sqrt{\alpha ^{}})^{p+1}g_s}}Str\sqrt{det(\eta _{mn}I+T^1F_{mn})}`$ (3.24)
where $`m,n=0,\mathrm{},5`$ and $`X`$ is in the direction one and $`i,j=2,\mathrm{},5`$ correspond to the spatial directions of the D4 brane. In action (3.24) we have we have reinstated the brane tension (i.e. string coupling and factors of $`\alpha ^{}`$). By expanding the DBI action and comparing with the Yang-Mills-Higgs actions one finds that the Yang-Mills coupling is related to the string parameters as
$`{\displaystyle \frac{1}{g_{YM}^2}}={\displaystyle \frac{T^{\frac{p+1}{2}}}{(2\pi \sqrt{\alpha ^{}})^{p+1}g_s}}.`$ (3.25)
Varying (3.24) with respect to $`A`$ gives
$`\delta L={\displaystyle \frac{1}{g_{YM}^2}}Str\left(\sqrt{det(TI+F)}[(TI+F)^1]^{mn}(D_m\delta A_nD_n\delta A_m)\right).`$ (3.26)
Minimising the expression for the energy density of the Yang-Mills-Higgs langrangian, the authors of have found the BPS equations for the dyonic instanton to be
$`E_i=TD_iX`$
$`F_{ij}=\stackrel{~}{F}_{ij}.`$ (3.27)
with $`i=2\mathrm{}5`$ and $`E_iF_{0i}`$. In addition to the BPS equations one has to satisfy the Gauss law
$`D_iE^i=T^2[X,D_0X].`$ (3.28)
After performing the variation, we are allowed to plug the BPS equations (3) into (3.26). For the BPS equations the square root in (3.26) linearises:
$$det(I+F)=\left(1\frac{1}{4}trF^2\right)^2$$
(3.29)
where $`tr`$ denotes the trace over the Lorentz indices. Going to the basis in which $`F`$ is skew-diagonal and evaluating the inverse of $`(I+F)^1`$ for the BPS configuration, one gets
$`(I+F)^1={\displaystyle \frac{1}{(1\frac{1}{4}trF^2)}}\left(\begin{array}{cccccc}(1\frac{1}{4}trF^2)& 0& 0& 0& 0& 0\\ 0& (1\frac{1}{4}trF^2)& 0& 0& 0& 0\\ 0& 0& 1& f& 0& 0\\ 0& 0& f& 1& 0& 0\\ 0& 0& 0& 0& 1& f\\ 0& 0& 0& 0& f& 1\end{array}\right),`$ (3.36)
where $`f^2=\frac{1}{4}trF^2`$. So we can rewrite (3.26) as
$`\delta L=Str\left(F^{ij}(D_i\delta A_jD_j\delta A_i)\right).`$ (3.37)
Since this expression now only involves two factors, there is no need to distinguish between $`Str`$ and ordinary $`tr`$. One can now apply ordinary integration by parts to obtain the equation of motion <sup>5</sup><sup>5</sup>5Note that the equations of motion for $`X`$ and $`A_0`$ are automatically satisfied when we plug the BPS conditions into (3.26).
$`D_iF^{ij}=0,`$ (3.38)
which is satisfied for the self-dual configurations due to the Bianchi identity.
The energy of the system is obtained by performing a Legendre transformation of the Lagrangian (2.14). Starting from the Lagrangian (2.14) and expanding the determinant under the square root one gets
$`L`$ $`=`$ $`{\displaystyle \frac{1}{g_{YM}^2}}Str[1+T^2DX.DXE^2+T(E.DX)^2T^2(DX.DX)E^2{\displaystyle \frac{1}{2}}F^2+T^2DX.F^2.DXE.F^2.E`$ (3.39)
$``$ $`{\displaystyle \frac{T^2}{2}}(DX.DX)trF^2+{\displaystyle \frac{1}{2}}E^2trF^2{\displaystyle \frac{1}{4}}trF^4+{\displaystyle \frac{1}{8}}(trF^2)^2]^{\frac{1}{2}}.`$
Next apply the Legendre transformation
$`=STr({\displaystyle \frac{L}{E}}.E)L.`$ (3.40)
The energy of the system is
$`={\displaystyle \frac{1}{g_{YM}^2}}Str((1`$ $`+`$ $`T^2DX.DX{\displaystyle \frac{1}{2}}F^2+T^2DX.F^2.DX{\displaystyle \frac{T^2}{2}}(DX.DX)trF^2`$ (3.41)
$`+`$ $`{\displaystyle \frac{1}{4}}trF^4+{\displaystyle \frac{1}{8}}(trF^2)^2)det(\mathrm{}.)^{\frac{1}{2}})`$
where here $`det(\mathrm{}.)`$ is short for the expression under the square root of (3.39). The denominator in (3.41) should be understood as the inverse of the hat function with respect to the star product. For the BPS configuration the energy density becomes
$`={\displaystyle \frac{1}{g_{YM}^2}}Str\left(1+E^2{\displaystyle \frac{1}{4}}trF^2\right).`$ (3.42)
This is the same expression as was obtained in by BPS arguments to minimize the energy. Note however, that the solution that we get is not obtained by minimizing the energy, but by directly showing that the BPS equations of solve the full equations of motion.
The second of the BPS equations (3) is the ordinary (anti)self duality equation for the nonabelian instanton. The solution in the singular gauge is given by
$`A_i=2{\displaystyle \frac{\rho ^2}{x^2(x^2+\rho ^2)}}\eta _{ij}^ax_j{\displaystyle \frac{\sigma ^a}{2}},`$ (3.43)
where $`\eta ^a`$ are the self-dual โt Hooft matrices, $`\sigma ^a`$ are the SU(2) generators, and the parameter $`\rho `$ is the size of the instanton. For the static field configuration ($`_0=0`$), after setting $`A_0=X`$, and after use of the Gauss law (3.28), the second BPS equation reduces to the covariant Laplace equation in the background of instanton
$`D^2X=0.`$ (3.44)
This equation has a unique solution for each $`v`$-VEV of the scalar field X. It is given by
$$X=\frac{g_{YM}}{T}v\frac{x^2}{(x^2+\rho ^2)}\frac{\sigma ^3}{2}.$$
(3.45)
When the BPS equation are satisfied, the field configuration has a mass which is equal to the sum of the charges
$`k`$ $`=`$ $`{\displaystyle \frac{1}{g_{YM}^2}}tr{\displaystyle _{WS}}d^4xF_{ij}F^{ij},`$ (3.46)
$`q`$ $`=`$ $`{\displaystyle \frac{T}{vg_{YM}^2}}tr{\displaystyle _{WS}}d^4xEDX={\displaystyle \frac{T}{vg_{YM}^2}}tr{\displaystyle _{S_{\mathrm{}}^3}}๐\stackrel{}{\sigma }.\stackrel{}{E}X=4\pi ^2\rho ^2v,`$ (3.47)
where $`WS`$ stands for world space of the D4 brane, and in the second equation when going to the surface integral we have used Gaussโ law. Charge $`k`$ is proportional to the instanton number, while $`Q=qv`$ is total electric charge carried by the instanton. Since $`Q`$ is a conserved charge we see that instanton stabilises at size $`\rho =\frac{Q^{1/2}}{2\pi v}`$ which is determined by the electric charge of the instanton and the VEV of the scalar field. Note that the electric charge (i.e. the electric part of the energy) is independent of the string coupling while the magnetic part depends on the Yang-Mills coupling as $`\frac{1}{g_{YM}^2}`$ i.e. as $`\frac{1}{g_s}`$. This suggests that in the brane picture the electric part of the configuration originates from the fundamental brane (string) while the magnetic part corresponds to the D-brane. That this is indeed the case will be shown in the next section.
To conclude this section we shall comment briefly on the abelian solution. One can easily derive the abelian BPS equations by minimizing the expression for the energy. They are the same as (3) up to a replacement $`D`$. Because of this replacement, the equations for $`E`$ and $`F`$ are no longer coupled, as in the non-abelian case. The equation for $`E`$, after use of Gaussโ law, implies that $`X`$ is a harmonic function on the worldspace ($`X=q/4\pi r^2`$), while the selfduality equation has the solution
$$A_a=\underset{r=1}{\overset{r=3}{}}(I^r)_{ab}\delta ^{bc}_cV_r$$
(3.48)
where the $`I^r`$ are three independent, anti-self-dual, complex structures in $`\text{}^2`$ obeying the quaternionic algebra relations and $`V`$ is harmonic on the world space $`\text{}^2`$ ($`V=\frac{K}{r^2}`$). Both the solution for $`X`$ and the solution for $`A`$ are singular and of infinite energy. Following the logic of one can try to get a non-singular, finite energy solution by embedding the brane in the appropriate background. Presence of the brane background induces a โcut-offโ in the distances which one can probe on the worldvolume of the $`D4`$ brane, leading to a finite energy of the electric part of the solution ($`E_{electric}=\frac{T}{g_{YM}^2}d^4\sigma \stackrel{}{E}\stackrel{}{}X=qv`$). However, the part of the energy corresponding to the instanton behaves as $`\frac{T}{g_{YM}^2}_\text{}^4๐\sigma ^4F_{ab}\stackrel{~}{F}_{ab}_{S_{r=r_0}^3}AA(\frac{v}{q})^2K^2`$. We see that the magnetic part of the solution depends on the separation of the brane from the background (v) and moreover that this dependence is quadratic. We have not yet been able to understand the cause of this behaviour of the abelianised solution.
## 4 Brane interpretation of the solution
It is known that the instanton solution in the gauge theory living on the stack of $`D4`$ branes has an interpretation as a $`D0`$ brane within the $`D4`$ brane . A large instanton is interpreted as a $`D0`$ brane which is dissolved into the flux of the YM gauge field, while a singular instanton corresponds to an undissolved $`D0`$ brane within the $`D4`$ brane.
The reasons for this interpretation are the following. The number of supersymmetries in the field theory preserved by the instanton solution is 1/2 of the supersymmetries of the brane theory. This is the same as the number of supersymmetries preserved by the $`D0D4`$ system. Furthermore, the instanton solution carries the same RR charge as a $`D0`$ brane. This can be seen by looking at the Chern-Simons (CS) term in the DBI action of the $`D4`$ brane. In a constant $`C`$-field, the CS term with selfdual gauge field $`F`$ reduces to
$`{\displaystyle \frac{1}{2}}T^2T_4{\displaystyle C_1}Tr(FF)=T_0{\displaystyle C_1},`$ (4.49)
since $`Tr(FF)=8\pi ^2`$ and $`T_0=(2\pi \sqrt{\alpha ^{}})^4T_4`$. So we recover the coupling of the $`D0`$ brane to the RR field.
Finally, the moduli of the effective action describing a system of two $`D4`$ branes and one $`D0`$ brane in the Higgs branch are the same as the moduli of the $`SU(2)`$ instanton. To illustrate this fact lets consider the case of the $`k=1`$, $`SU(2)`$ instanton. The number of moduli of this instanton is eight: four corresponding to the position of the instanton, three corresponding to the orientation of the instanton in the $`SU(2)`$ group (i.e. the embedding of the $`U(1)`$ in the $`SU(2)`$ group) and one corresponding to the instanton size.
On the other side, to count the number of the moduli in the $`D0D4`$ system we have to look at the potential term in the effective action for the $`D0D4`$ system
$`{\displaystyle \frac{g_{D0}^2}{4}}{\displaystyle \underset{A=1}{\overset{3}{}}}(\chi _{ia}^{}\sigma _{ij}^A\chi _{ja})^2+{\displaystyle \underset{i=1}{\overset{5}{}}}(X_iX_i^{})^2\chi ^{}\chi .`$ (4.50)
Here $`\chi _i`$ is a doublet of complex hypermultiplet scalars, the index $`a=1,2`$ labels two D4 branes and, $`X_i`$ and $`X_i^{}`$ are scalars in the vector multiplet.
In the Higgs branch, the $`D0`$ brane is within the $`D4`$ brane ($`X_i=X_i^{}`$), so the second term in (4.50) vanishes identically. Each of the remaining three terms in (4.50) has to vanish separately, and so we have three constraints on the field $`\chi `$. Moreover, using the $`SU(2)`$ invariance of (4.50) we can fix one component of $`\chi `$. There are eight real hypermultiplet scalars in $`\chi _a`$. After use of the constraints and $`SU(2)`$ invariance one is left with four moduli. These moduli correspond to moduli describing the embedding of the instanton and instanton size. The additional four moduli come from the $`D0D0`$ strings and describe the position of the $`D0`$ brane within the $`D4`$ brane.
The instanton solution in our case is unmodified compared to the un-Higgsed YM theory, except that the size of the instanton is no longer a free modulus, but it is fixed by the separation of the branes and by the electric charge which the instanton carries. Hence, we can interpret our instanton as a $`D0`$ brane dissolved in the YM flux in the worldvolumes of two separated $`D4`$ branes.
However, we still have to explain the origin of the electric field on the D4 brane. A useful observation is that the electric part of the energy of the system is independent of the string coupling. This is a know property of the energy of fundamental objects, such as $`F1`$ strings. This is in contrast to the magnetic part of the energy which is proportional to $`\frac{1}{g_s}`$, i.e. it behaves as the mass of the D-brane.
Additional information is gained from the effective geometry of the configuration. So let us look more closely at the shape of the brane for our field configuration.
Along the lines of one can immediately read off the shape of the brane from (3.45). This is depicted in the fig 1.a. Far away from the core of the instanton, the shape of the brane looks like the brane pulled by fundamental strings, i.e. $`X`$ behaves as $`X\frac{v}{2}(1\frac{\rho ^2}{r^2})`$. However, at $`r_0=\frac{\rho }{\sqrt{3}}`$ the brane changes shape from concave to convex. Also, if one calculates the energy density corresponding to the electric field, i.e. the intensity of the electric field (see fig 1.b), one finds the expression
$$_E=E_i^2\frac{r^2}{(r^2+\rho ^2)^4}.$$
(4.51)
This expression has a maximum at distance $`r_0=\frac{\rho }{\sqrt{3}}`$, at the same position at which the shape of the brane changes from concave to convex. Hence, it seems like the source of the electric field on the brane is located on the sphere of radius $`r_0`$. To determine the origin of the electric field we first recall that the D-brane is perturbatively defined in terms of $`F1`$ strings with Chan-Paton (CP) factors attached to the end of the string. CP factors are sources of the gauge group on the brane and different brane configurations in the effective description correspond to different gauge configurations. For example, an $`F1`$ string ending on the $`D3`$ brane in the effective description is an electric charge in the world volume gauge theory of the $`D3`$ brane , and it turns on the electric field in the worldvolume of the brane. On the other hand a $`D1`$ string turns on the magnetic components of the gauge field in the worldvolume of the brane, and is represented as a magnetic monopole in the worldvolume gauge theory.
When the $`F1`$ (or ($`p,q`$) string) terminates on the brane, the effective geometry of the brane is such that it bends under the tension of the string. Due to the fact that the brane has a tendency to minimize its surface area, the brane has a convex form when pulled by the string. In fig 2. the shape of two separated branes, pulled by the fundamental string which connects them, is shown. For the configuration shown in fig 2. the electric field on the branes grows as we approach the point at which the string is attached.
In our case however, the effective geometry is different since the source of the electric field is not localised at a point.
Far from the core of the instanton, the electric field falls-off as $`\frac{1}{x^3}`$. This is exactly the behaviour of the electric charge in 4+1 dimensions. So, far away from the core of the instanton, the geometry of the brane is the same as in the previous case where the $`F1`$ pulled the brane. All this suggests that the correct interpretation of the electric field on the brane is that it originates from the $`F1`$ string which connects $`D0`$ and $`D4`$ branes. The effective description of a $`D0`$ brane within a $`D4`$ brane is that of the instanton: in other words when the $`D0`$ brane is in the $`D4`$ brane it dissolves into the flux. Hence, since the $`F1`$ string terminates on the $`D0`$, the same thing happens to the $`F1`$ string: it dissolves into the electric flux on the brane. As we start to separate the branes, the instanton starts to shrink, but its complete collapse is halted by the electrostatic repulsion (which acts as negative tension on the brane surface). The presence of the delocalised electric source causes the โblowing upโ of the brane surface which counterbalances the brane which tends to collapse. <sup>6</sup><sup>6</sup>6This is to some extend similar to the Dirac model of the electron as a charged bubble, which is prevented from collapse by electric repulsion, except that in our case the electric field is confined to live just on the surface of the brane and not in the full spacetime. As we increase the separation between the branes, the force that pulls the brane, and hence the $`D0`$ within it, increases <sup>7</sup><sup>7</sup>7The tension of the string connecting the separating branes is constant, but the total force acting on the branes is $`vT`$, so it increases as we separate the branes. and hence the radius at which the instanton stabilises becomes smaller.
In summary, if we start with the system of coinciding $`D4`$ branes and a $`D0`$ brane within them, then the presence of the $`D0`$ brane in the effective theory is seen as either a โlargeโ $`\rho `$ instanton (a dissolved $`D0`$ brane, which is described by the Higgs branch of the effective theory) or a zero size instanton. The singular instanton is interpreted as the $`D0`$ brane that can be pulled out from the stack of $`D4`$ branes. Once we separate the $`D0`$ brane from the stack of the $`D4`$ branes, the system is described by the Coulomb branch of the full effective action (4.50). However, if we separate the stack of $`D4`$ branes with the โlargeโ $`D0`$ brane (instanton) sitting within them, then the $`D0`$ brane will start shrinking and it will stabilise at a radius which is inversely proportional to the separation between the branes. <sup>8</sup><sup>8</sup>8In the previous discussion we were talking about BPS configurations, in which kinetic energy of the system was always zero. Hence, when we talk about the process of separating the branes, or shrinking of the instanton, we assume a quasi-static process between two BPS configurations.
It is interesting to note <sup>9</sup><sup>9</sup>9I thank P.K.Townsend for pointing out this to me. that for each dyonic instanton (characterised by charge $`Q=qv`$ and instanton number $`k`$), there is a corresponding configuration of a string of length $`v`$ and separated set of $`kD0`$ branes which has the same energy as the dyonic instanton. However this state cannot be described in the field theory living on the $`D4`$ brane. Although these two configurations have the same energy, it seems likely that the regions of the full moduli space, corresponding to the dyonic instanton and the $`D0F1`$ system, are separated by a potential barrier. In support of this, we note that to โpull outโ a $`D0`$ brane one first has to shrink it to zero size, and that requires infinite separation of the $`D4`$ branes. However, to describe the full moduli space one would need to analyse the effective action of the full $`D0D4`$ system which we leave for future investigation.
## 5 Conclusions and discussion
In the first part of the paper we have tried to present systematic rules that one should use when dealing with the nonabelian DBI action. These rules allows one to treat the NDBI action as abelian, and work in closed form (with all terms in the expansion). However, we have pointed out that there are limitations to this approach, i.e that after certain operations (like taking covariant derivatives of hat functions) one loses the ability to further express things in a obvious closed form. As for future directions, it would be interesting to try to apply this approach to deduce the Hamiltonian formulation of the NDBI action. In this case one is facing the new problem of solving the nonabelian constraints. We hope to address this question in future work.
In the second part we have seen a new effect in the effective brane geometryโ the blowing up of the brane. It would we interesting to try to use this effect to construct compact branes. Up to now all compact brane configurations were constructed either by wrapping branes on a compact submanifold of the background (in this case the brane is prevented from collapse by the background geometry) or by putting the brane in an external RR-field . It would be interesting to explore the possibility of having compact branes due to a nontrivial gauge configuration on the worldvolume of the brane.
## Acknowledgements
I would like to thank Dominic Brecher, Neil Constable, David Tong, Ian Hawke, Eduardo Eyras, Nick Manton, Jonathan Moore, Malcolm Perry, Andy Neitzke, Adam Ritz for discussions. I am especially grateful to Paul Townsend and Kasper Peeters for many useful discussions, a lot of patience and great moral support.
|
warning/0006/astro-ph0006115.html
|
ar5iv
|
text
|
# Wide Field Imaging. I. Applications of Neural Networks to object detection and star/galaxy classification
## 1 Introduction
Astronomical wide field (hereafter WF) imaging encompasses the use of images larger than $`4000^2`$ pxls (Lipovestky 1993) and is the only tool to tackle problems based on rare objects or on statistically significant samples of optically selected objects. Therefore, WF imaging has been and still is of paramount relevance to almost all field of astrophysics: from the structure and dynamics of our Galaxy, to the environmental effects on galaxy formation and evolution, to the large scale structure of the Universe. In the past, WF was the almost exclusive domain of Schmidt telescopes equipped with large photographic plates and was the main source of targets for photometric and spectroscopic follow-upโs at telescopes in the 4 meter class. Nowadays, the exploitation of the new generation 8 meter class telescopes, which are designed to observe targets which are often too faint to be even detected on photographic material (the POSS-II detection limit in B is $`21.5`$ $`mag)`$, requires digitised surveys realized with large format CCD detectors mounted on dedicated telescopes. Much effort has therefore been devoted worldwide to construct such facilities: the MEGACAM project at the CFH, the ESO Wide Field Imager at the 2.2 meter telescope, the Sloan - DSS and the ESO-OAC VST (Mancini et al. 1999) being only a few among the ongoing or planned experiments.
One aspect which is never too often stressed is the humongous problem posed by the handling, processing and archiving of the data produced by these instruments: the VST alone, for instance, is expected to produce a flow of almost 30 GByte of data per night or more than 10 Tbyte per year of operation.
The scientific exploitation of such a huge amount of data calls for new data reduction tools which must be reliable, must require a small amount of interactions with the operators and need to be as much independent on a priori choices as possible.
In processing a WF image, the final goal is usually the construction of a catalogue containing as many as possible astrometric, geometric, morphological and photometric parameters for each individual object present on the image. The first step in any catalogue construction is therefore the detection of the objects, a step which, as soon as the quality of the images increases (both in depth and in resolution), becomes much less obvious than what it may seem at first glance. The traditional definition of โobjectโ as a set of connected pixels having brightness higher than a given threshold, has in fact several well known pitfalls. For instance, low surface brightness galaxies very often escape recognition since i) their central brightness is often comparable or fainter than the detection threshold, and ii) their shape is clumpy, which implies that even though there may be several nearby pixels above the threshold, they can often be not connected and thus escape the assumed definition.
A similar problem is also encountered in the catalogues extracted from the Hubble Deep Field (HDF) where a pletora of small โclumpyโ objects is detected but it is not clear whether each clump represents an individual object or rather is a fragment of a larger one. Ferguson (1998) stresses some even stronger limitations of the traditional approach to object detection: i) a comparison of catalogues obtained by different groups from the same raw material and using the same software shows that, near the detection limits the results are strongly dependent on the assumed definition of โobjectโ; ii) object detection performed by the different groups is worse than what even an untrained astronomer can attain by visually inspecting an image: many objects which are present in the image are lost by the software, while others which are missing on the image are detected (hereafter spurious objects; see Fig. 1). In other words: the silent assumption that faint objects consist of connected and amorphous sets of pixels makes the definition of โobjectโ astronomerโdependent and produces quite ambiguous results at very low S/N ratios.
The classification on morphological grounds only of an object as a star or as a galaxy substantially relies on whether the object is spatially resolved or not. Human experts can usually classify objects either directly from the appearance of the objects on an image (either photographic or digital) or from the value of some finite set of derived parameters via diagnostic diagrams (such as magnitude versus area). This approach, however, is too much time consuming and too much dependent on the โknow howโ and personal experience of the observer: i) the choice of the most suitable parameters largely varies from author to author thus making comparisons difficult if not at all impossible, and ii) regardless the complexity of the problem, due to the obvious limitations of representing three or more dimensions spaces on a two dimensional graph, only two features are often considered. In recent years much effort has therefore been devoted to implement and fine tune Artificial Intelligence (herafter AI) tools to perform star/galaxy classification on authomatic grounds. The powerful package SExtractor (SEx, Bertin & Arnouts 1996), for instance, uses nine features (eight isophotal areas and the peak intensity) and a neural network to classify objects. The SEx output is an index, ranging from 0 to 1, which gives the degree of โstellarityโ of the object. This implies, however, still a fair degree of arbitrarity in choosing these features and not any other set. Other approaches to the same problems will be reviewed in the last section of this paper.
This paper is divided in two major parts: in ยง2 we present the AI theory used in the paper, and in ยง3 the experiments. Finally, in ยง4 we discuss our results and draw the conclusions.
## 2 The Theory
In the AI domain there are dozens of different NNโs used and optimised to perform the most various tasks. In the astronomical literature, instead, only two types of NNโs are used: the ANN, called in the AI literature Multi-layer Perceptron (MLP) with back-propagation learning algorithm, and the Kohonenโs Self-organizing Maps (or their supervised generalization).
We followed a rather complex approach which can be summarised as follows: Principal Component Analysis (PCA) NNโs were used to reduce the dimensionality of the input space. Supervised NNโs need a large amount of labelled data to obtain a good classification while unsupervised NNโs overcome this need, but do not provide good performances when classes are not well separated. Hybrid and unsupervised hierarchical NNโs are therefore very often introduced to simplify the expensive post-processing step of labelling the output neurons in classes (such as objects/background), in the object detection process. In the following subsections we illustrate the properties of several types of NNโs which were used in one or another of the various tasks. All the discussed models were implemented, trained and tested and the results of the best performing ones are illustrated in detail in the next sections.
### 2.1 PCA Neural Nets
A pattern can be represented as a point in a $`L`$-dimensional parameter space. To simplify the computations, it is often needed a more compact description, where each pattern is described by $`M`$, with $`M<L`$, parameters. Each $`L`$-dimensional vector can be written as a linear combination of $`L`$ orthonormal vectors or as a smaller number of orthonormal vectors plus a residual error. PCA is used to select the orthonormal basis which minimizes the residual error.
Let $`x`$ be the $`L`$-dimensional zero mean input data vectors and $`C=E(xx^T)=xx^T`$ be the covariance matrix of the input vectors $`x`$. The $`i`$-th principal component of $`x`$ is defined as $`x^Tc(i)`$, where $`c(i)`$ is the normalized eigenvector of $`C`$ corresponding to the $`i`$-th largest eigenvalue $`\lambda (i)`$.
The subspace spanned by the principal eigenvectors $`c(1),\mathrm{},c(M),(M<L)`$ is called the PCA subspace (with dimensionality $`M`$; Oja 1982; Oja et al. 1996). In order to perform PCA, in some cases and expecially in the non linear one, it is convenient to use NNโs which can be implemented in various ways (Baldi & Hornik 1989; Jutten & Herault 1991; Oja 1982; Oja, Ogawa & Wangviwattana 1991; Plumbley 1993; Sanger 1989). The PCA NN used by us was a feedforward neural network with only one layer which is able to extract the principal components of the stream of input vectors. Fig. 2 summarises the structure of the PCA NNโs. As it can be seen, there is one input layer, and one forward layer of neurons which is totally connected to the inputs. During the learning phase there are feedback links among neurons, the topology of which classifies the network structure as either hierarchical or symmetric depending on the feedback connections of the output layer neurons.
Typically, Hebbian type learning rules are used, based on the one unit learning algorithm originally proposed in (Oja 1982). The adaptation step of the learning algorithm - in this case the network is composed by only one output neuron - is then written as:
$$w_j^{(t+1)}=w_j^{(t)}+\mu y^{(t)}\left[x_j^{(t)}y^{(t)}w_j^{(t)}\right]$$
(1)
where $`x_j^{(t)}`$, $`w_j(t)`$ and $`y^{(t)}`$ are, respectively, the value of the $`j`$-th input, of the $`j`$-th weight and of the network output at time $`t`$, while $`\mu `$ is the learning rate. $`\mu y^{(t)}x_j^{(t)}`$ is the Hebbian increment and Eq.1 satisfies the condition:
$$\underset{j=1}{\overset{M}{}}\left(w_j^{(t)}\right)^21$$
(2)
Many different versions and extensions of this basic algorithm have been proposed in recent years (Karhunen & Joutsensalo 1994 = KJ94; Karhunen & Joutsensalo 1995 = KJ95; Oja et al. 1996; Sanger 1989).
The extension from one to more output neurons and to the hierarchical case gives the well known Generalized Hebbian Algorithm (GHA) (Sanger 1989; KJ95):
$$w_{ji}^{(t+1)}=w_{ji}^{(t)}+\mu y_j^{(t)}\left[x_i^{(t)}\underset{k=1}{\overset{j}{}}y_k^{(t)}w_{jk}^{(t)}\right]$$
(3)
while the extension to the symmetric case gives the Ojaโs Subspace Network (Oja 1982):
$$w_{ji}^{(t+1)}=w_{ji}^{(t)}+\mu y_j^{(t)}\left[x_i^{(t)}\underset{k=1}{\overset{M}{}}y_k^{(t)}w_{jk}^{(t)}\right]$$
(4)
In both cases the weight vectors must be orthonormalized and the algorithm stops when:
$$t_n=\sqrt{\underset{j=1}{\overset{M}{}}\underset{i=1}{\overset{L}{}}\left(w_{ji}^{(t)}w_{ji}^{(t1)}\right)^2}<\epsilon $$
where $`\epsilon `$ is an arbitrarily choosen small value. After the learning phase, the network becomes purely feedforward. KJ94 and KJ95 proved that PCA neural algorithms can be derived from optimization problems, such as variance maximization and representation error minimization. They generalized these problems to nonlinear problems, deriving nonlinear algorithms (and the relative networks) having the same structure of the linear ones: either hierarchical or symmetric. These learning algorithms can be further classified in robust PCA algorithms and nonlinear PCA algorithms. KJ95 defined robust PCA as those in which the objective function grows slower than a quadratic one. The non linear learning function appears at selected places only. In nonlinear PCA algorithms all the outputs of the neurons are nonlinear function of the responses.
More precisely, in the robust generalization of variance maximization, the objective function $`f(z)`$ is assumed to be a valid cost function such as $`lncos(z)`$ or $`|z|`$. This leads to the adaptation step of the learning algorithm:
$$w_{ji}^{(t+1)}=w_{ji}^{(t)}+\mu g\left(y_j^{(t)}\right)e_{ji}^{(t)}$$
(5)
where:
$$e_{ji}^{(t)}=x_i^{(t)}\underset{k=1}{\overset{l(j)}{}}y_k^{(t)}w_{jk}^{(t)}$$
$$g=\frac{df}{dz}$$
In the hierarchical case $`l(j)=j`$. In the symmetric case $`l(j)=M`$, the error vector $`e_j^{(t)}`$ becomes the same $`e^{(t)}`$ for all the neurons, and Eq. 5 can be compactly written as:
$`W^{(t+1)}`$ $`=`$ $`W^{(t)}+\mu \left(IW^{(t)}W^{(t)T}\right)xg\left(x^TW^{(t)}\right)`$
$`=`$ $`W^{(t)}+\mu e^{(t)}g\left(y^{(t)T}\right)`$
where $`y^{(t)T}=x^TW^{(t)}`$ is the instantaneous vector of neuron responses at time $`t`$. The learning function $`g`$, derivative of $`f`$, is applied separately to each component of the argument vector.
The robust generalisation of the representation error problem (KJ95) with $`f(t)t^2`$, leads to the stochastic gradient algorithm:
$$w_j^{(t+1)}=w_j^{(t)}+\mu \left(w_j^{(t)T}g\left(e_j^{(t)}\right)x^{(t)}+x^{(t)T}w_j^{(t)}g\left(e_j^{(t)}\right)\right)$$
(7)
This algorithm can be again considered in both the hierarchical and symmetric cases. In the symmetric case $`l(j)=M`$, the error vector is the same $`e^{(t)}`$ for all the weights $`w^{(t)}`$. In the hierarchical case $`l(j)=j`$, Eq. 7 gives the robust counterparts of principal eigenvectors $`c(i)`$.
In Eq. 7 the first update term
$$w_j^{(t)T}g\left(e_j^{(t)}\right)x^{(t)}$$
is proportional to the same vector $`x(t)`$ for all weights $`w_j(t)`$. Furthermore, we can assume that the error vector $`e(t)`$ is relatively small after the initial convergence. Hence, we can neglect the first term in Eq. 7 and this leads to:
$$w_j^{(t+1)}=w_j^{(t)}+\mu y_j^{(t)}g\left(e_j^{(t)}\right)$$
(8)
Let us consider now the nonlinear extensions of PCA algorithms which can be obtained in a heuristic way by requiring all neuron outputs to be always nonlinear in Eq. 5, then:
$$w_j^{(t+1)}=w_j^{(t)}+\mu g\left(y_j^{(t)}\right)b_j^{(t)}$$
(9)
where:
$$b_j^{(t)}=x^{(t)}\underset{k=1}{\overset{l(j)}{}}g\left(y_k^{(t)}\right)w_k^{(t)}$$
In previous experiments (Tagliaferri et al. 1999, Tagliaferri et al. 1998) we found that the hierarchical robust NN of Eq. 5 with learning function $`g=tanh(\alpha x)`$ achieves the best performance with respect to all the other mentioned PCA NNโs and linear PCA.
### 2.2 Unsupervised neural nets
Unsupervised NNโs partition the input space into clusters and assign to each neuron a weight vector which univocally individuates the template characteristic of one cluster in the input feature space. After the learning phase, all the input patterns are classified.
Kohonen (1982, 1988) Self Organizing Maps (SOM) are composed by one neuron layer structured in a rectangular grid of $`m`$ neurons. When a pattern $`x`$ is presented to the NN, each neuron $`i`$ receives the input and computes the distance $`d_i`$ between its weight vector $`w_i`$ and $`x`$. The neuron which has the minimum $`d_i`$ is the winner. The adaptation step consists in modifying the weights of the neurons in the following way:
$$w_j^{(t+1)}=w_j^{(t)}+\epsilon ^{(t)}h_{\sigma ^{(t)}}\left(d(j,k)\right)\left(xw_j^{(t)}\right)$$
(10)
where $`\epsilon ^{(t)}`$ is the learning rate ($`0\epsilon ^{(t)}1`$) decreasing in time, $`d(j,k)`$ is the distance in the grid between the $`j`$ and the $`k`$ neurons and $`h_{\sigma ^{(t)}}\left(x\right)`$ is a unimodal function with variance $`\sigma ^{\left(t\right)}`$ decreasing with $`x`$.
The Neural-Gas NN is composed by a linear layer of neurons and a modified learning algorithm (Martinetz Berkovitch & Shulten 1993). It classifies the neurons in an ordered list $`(j_1,\mathrm{},j_m)`$ accordingly to their distance from the input pattern. The weight adaptation depends on the position $`rank(j)`$ of the $`j`$-th neuron in the list:
$$w_j^{(t+1)}=w_j^{(t)}+\epsilon ^{(t)}h_{\sigma ^{(t)}}\left(rank\left(j\right)\right)\left(xw_j^{(t)}\right)$$
(11)
and works better than the preceding one: in fact, it is quicker and reaches a lower average distortion value<sup>1</sup><sup>1</sup>1Let $`P(x)`$ be the pattern probability distribution over the set $`V\mathrm{}^n`$ and let $`w_i(x)`$ be the weight vector of the neuron which classifies the pattern $`x`$. The average distortion is defined as $`E=P(x)\left(xw_i(x)\right)^2๐x`$.
The Growing Cell Structure (GCS) (Fritzke 1994) is a NN which is capable to change its structure depending on the data set. Aim of the net is to map the input pattern space into a two-dimensional discrete structure $`S`$ in such a way that similar patterns are represented by topological neighboring elements. The structure $`S`$ is a two-dimensional simplex where the vertices are the neurons and the edges attain the topological information. Every modification of the net always maintains the simplex properties. The learning algorithm starts with a simple three node simplex and tries to obtain an optimal network by a controlled growing process: id est, for each pattern $`x`$ of the training set, the winner $`k`$ and the neighbors weights are adapted as follows:
$$w_k^{(t+1)}=w_k^{(t)}+\epsilon _b\left(xw_k^{(t)}\right);w_j^{(t+1)}=w_j^{(t)}+\epsilon _n\left(xw_j^{(t)}\right)$$
(12)
$`j`$ connected to $`k`$; where $`\epsilon _b`$ and $`\epsilon _n`$ are constants which determine the adaptation strength for the winner and for the neighbors, respectively.
The insertion of a new node is made after a fixed number $`\lambda `$ of adaptation steps. The new neuron is inserted between the most frequent winner neuron and the more distant of its topological neighbors. The algorithm stops when the network reaches a pre-defined number of elements.
The on-line K-means clustering algorithm (Lloyd 1982) is a simpler algorithm which applies the Gradient Descent (=GD) directly to the average distortion function as follows:
$$w_j^{(t+1)}=w_j^{(t)}+\epsilon ^{(t)}\left(xw_j^{(t)}\right)$$
(13)
The main limitation of this technique is that the error function presents many local minima which stop the learning before reaching the optimal configuration.
Finally, the Maximum Entropy NN (Rose, Gurewitz & Fox, 1990) applies the GD to the error function to obtain the adaptation step:
$$w_j^{(t+1)}=w_j^{(t)}+\epsilon ^{(t)}\frac{\mathrm{exp}\left(\beta ^{(t)}d_j\right)}{_{k=1}^m\mathrm{exp}\left(\beta ^{(t)}d_k\right)}\left(xw_j^{(t)}\right)$$
(14)
where $`\beta `$ is the inverse temperature and takes value increasing in time and $`d_j`$ is the distance between the $`j`$-th and the winner neurons.
### 2.3 Hybrid neural nets
Hybrid NNโs are composed by a clustering algorithm which makes use of the information derived by one unsupervised single layer NN. After the learning phase of the NN, the clustering algorithm splits the output neurons in a number of subsets which is equal to the number of the desired output classes. Since the aim is to put similar input patterns in the same class and dissimilar input patterns in different classes, a good strategy consists in applying a clustering algorithm directly to the weight vectors of the unsupervised NN.
A non-neural agglomeration clustering algorithm that divides the pattern set (in this case the weights of the neurons) $`W=\{w_1,\mathrm{},w_m\}`$ in $`l`$ clusters (with $`l<m`$) can be briefly summarized as follows:
1. it initially divides $`W`$ in $`m`$ clusters $`C_1,\mathrm{},C_m`$ such that $`C_j=\left\{w_j\right\}`$;
2. then it computes the distance matrix $`D`$ with elements $`D_{ij}=d(C_i,C_j)`$;
3. then it finds the smallest element $`D_{ij}`$ and unifies the clusters $`C_i`$ and $`C_j`$ in a new one $`C_{ij}=C_iC_j`$;
4. if the number of clusters is greater than $`l`$ then it goes to step 2 else, it finally stops.
Many algorithms quoted in literature (Everitt 1977) differ only in the way in which the distance function is computed. For example:
$$d(C_i,C_j)=\underset{w_{ik}C_iandw_{jl}C_j}{\mathrm{min}}w_{ik}w_{jl}$$
(nearest neighbor algorithm);
$$d(C_i,C_j)=\frac{1}{\left|C_i\right|}\underset{w_{ik}C_i}{}w_{ik}\frac{1}{\left|C_j\right|}\underset{w_{jl}C_j}{}w_{jl}$$
(centroid method);
$$d(C_i,C_j)=\frac{1}{\left|C_i\right|\left|C_j\right|}\underset{w_{ik}C_i,w_{jl}C_j}{}w_{ik}w_{jl}$$
(average between groups).
The output of the clustering algorithm will be a labelling of the patterns (in this case neurons) in $`l`$ different classes.
### 2.4 Unsupervised hierarchical neural nets
Unsupervised hierarchical NNโs add one or more unsupervised single layers NN to any unsupervised NN, instead of a clustering algorithm as it happens in hybrid NNโs.
In this way, the second layer NN learns from the weights of the first layer NN and clusters the neurons on the basis of a similarity measure or a distance. The iteration of this process to a few layers gives the unsupervised hierarchical NNโs.
The number of neurons at each layer decreases from the first to the output layer and, as a consequence, the NN takes the pyramidal aspect shown in Fig. 3. The NN takes as input a pattern $`x`$ and then the first layer finds the winner neuron. The second layer takes the first layer winner weight vector as input and finds the second layer winner neuron and so on up to the top layer. The activation value of the output layer neurons is 1 for the winner unit and 0 for all the others. In short: the learning steps of a $`s`$ layer hierarchical NN with training set $`X`$ are the following:
* the first layer is trained on the patterns of $`X`$ with one of the learning algorithms for unsupervised NNโs;
* the second layer is trained on the elements of the set $`X_2`$ which is composed by the weight vectors of the first layer winner units;
* the process is iterated to the $`ith`$ layer NN ($`i>2`$) on the training set which is composed by the weight vectors of the winner neurons of the $`(i1)th`$ layer when presenting $`X`$ to the first layer NN, $`X_2`$ to the second layer and so on.
By varying the learning algorithms we obtain different NNโs with different properties and abilities. For instance, by using only SOMโs we have a Multi-layer SOM (ML-SOM) (Koh J., Suk & Bhandarkar 1995) where every layer is a two-dimensional grid. We can easily obtain (Tagliaferri, Capuano & Gargiulo 1999) ML-NeuralGas, ML-Maximum Entropy or ML-K means organized on a hierarchy of linear layers. The ML-GCS has a more complex architecture and has at least 3 units for layer.
By varying the learning algorithms in the different layers, we can take advantage from the properties of each model (for instance, since we cannot have a ML-GCS with 2 output units we can use another NN in the output layer).
A hierarchical NN with a number of output layer neurons equal to the number of the output classes simplifies the expensive post-processing step of labelling the output neurons in classes, without reducing the generalization capacity of the NN.
### 2.5 Multi-layer Perceptron
A Multi-Layer Perceptron (MLP) is a layered NN composed by:
* one input layer of neurons which transmit the input patterns to the first hidden layer;
* one or more hidden layers with units computing a nonlinear function of their inputs;
* and one output layer with elements calculating a linear or a nonlinear function of their inputs.
Aim of the network is to minimize an error function which generally is the sum of squares of the difference between the desired output (target) and the output of the NN. The learning algorithm is called back-propagation since the error is back-propagated in the previous layers of the NN in order to change the weights. In formulae, let $`x^p`$ be an $`L`$dimensional input vector with corresponding target output $`c^p`$. The error function is defined as follows:
$$E_p=\frac{1}{2}\underset{i}{}\left(c_i^py_i^p\right)^2$$
where $`y_i^p`$ is the output of the $`ith`$ output neuron. The learning algorithm updates the weights by using the gradient descent (GD) of the error function with respect to the weights. If we define the input and the output of the neuron $`j`$ respectively as:
$$net_j^p=\underset{i}{}w_{ji}y_i^p$$
and
$$y_j^p=f\left(net_j^p\right)$$
where $`w_{ji}`$ is the connection weight from the neuron $`i`$ to the neuron $`j`$, and $`f\left(x\right)`$ is linear or sigmoidal for the output nodes and sigmoidal for the hidden nodes. It is well known in literature (Bishop 1995) that these facts lead to the following adaptation steps:
$$w_{ji}^{(t+1)}=w_{ji}^{(t)}+\mathrm{}w_{ji}^{(t)}\mathrm{where}\mathrm{}w_{ji}^{(t)}=\eta \delta _j^py_i^p$$
(15)
and
$$\delta _j^p=\left(c_j^py_j^p\right)y_j^p\left(1y_j^p\right)\mathrm{or}\delta _j^p=\left(c_j^py_j^p\right)\underset{k}{}\delta _k^pw_{kj}$$
(16)
for the output and hidden units, respectively. The value of the learning rate $`\eta `$ is small and causes a slow convergence of the algorithm. A simple technique often used to improve it is to sum a momentum term to Eq. 15 which becomes:
$$\mathrm{}w_{ji}^{(t)}=\eta \delta _j^py_i^p+\mu \mathrm{}w_{ji}^{(t1)}$$
(17)
This technique generally leads to a significant improvement in the performances of GD algorithms but it introduces a new parameter $`\mu `$ which has to be empirically chosen and tuned.
Bishop (1995) and Press et al. (1993) summarize several methods to overcome the problems related to the local minima and to the slow time convergence of the above algorithm. In a preliminary step of our experiments, we tried all the algorithms discussed in chapter 7 of Bishop (1995) finding that a hybrid algorithm based on the scaled conjugate gradient for the first steps and on the Newton method for the next ones, gives the best results with respect to both computing time and relative number of errors. In this paper we used it in the MLPโs experiments.
## 3 The Experiments
### 3.1 The data
In this work we use a 2000x2000 arcsec<sup>2</sup> area centered on the North Galactic Pole extracted from the slightly compressed POSS-II F plate n. 443, available via network at the Canadian Astronomy Data Center (http://cadcwww.dao.nrc.ca). POSS-II data were linearised using the sensitometric spots recorded on the plate. The seeing FWHM of our data was 3 arcsec. The same area has been widely studied by others and, in particular, by Infante & Pritchet (1992, =IP92) and Infante, Pritchet & Hertling (1995) who used deep observations obtained at the 3.6 m CFHT telescope in the $`F`$ photographic band under good seeing conditions (FWHM $`<1`$ arcsec) to derive a catalogue of objects complete down to $`m_F23`$, id est, much deeper than the completeness limit of our plate. Their catalogue is therefore based on data of much better quality and accuracy than ours, and it was for the availability of such good template that we decided to use this region for our experiments. We also studied a second region in the Coma cluster (which happens to be in the same n. 443 plate) but since none of the catalogues available in literature is much better than our data, we were forced to neglect it in most of the following discussion.
The characteristics of the selected region, a relatively empty one, slightly penalise our NN detection algorithms which can easily recognise objects of quite different sizes. On the contrary of what happens to other algorithms NExt works well even on areas where both very large and very small objects are present such as, for instance, the centers of nearby clusters of galaxies as our preliminary test on a portion of the Coma clusters clearly shows (Tagliaferri et al. 1998).
### 3.2 Structure of NExt
The detection and classification of the objects are a multi-step task:
1) First of all, following a widely used AI approach, we mathematically transform the detection task into a classification one by compressing the redundant information contained in nearby pixels by means of a nonโlinear PCA NNโs. Principal vectors of the PCA are computed by the NN on a portion of the whole image. The values of the pixels in the transformed $`M`$ dimensional eigen-space obtained via the principal vectors of the PCA NN are then used as inputs to unsupervised NNโs to classify pixels in few classes. We wish to stress that, in this step, we are still classifying pixels, and not objects.
The adopted NN is unsupervised, i.e. we never feed into the detection algorithm any a priori definition of what an object is, and we leave it free to find its own object definition. It turns out that image pixels are split in few classes, one coincident with what astronomers call background and some others for the objects (in the astronomical sense). Afterwords, the class containing the background pixels is kept separated from the other classes which are instead merged together. Therefore, as final output, the pixels in the image are divided in โobjectโ or โbackgroundโ.
2) Since objects are seldom isolated in the sky, we need a method to recognise overlapping objects and deblend them. We adopt a generalisation of the method used by Focas (Jarvis & Tyson 1981).
3) Due to the noise, object edges are quite irregular. We therefore apply a contour regularisation to the edges of the objects in order to improve the following star/galaxy classification step.
4) We define and measure the features used, or suitable, for the star/galaxy classification, then we choose the best performing features for the classification step, through the sequential backward elimination strategy (Bishop 1995).
5) We then use a subset of the IP92 catalog to learn, validate and test the classification performed by NExt on our images. The training set was used to train the NN, while the validation was used for model selection, i.e. to select the most performing parameters using an independent data set. As template classifier, we used SEx, whose classifier is also based on NNs.
The detection and classification performances of our algorithm were then compared with those of traditional algorithms, such as SEx. We wish to stress that in both the detection and classification phases, we were not interested in knowing how well NExt can reproduce SEx or the astronomerโs eye performances, but rather to see whether the SEx and NExt catalogs are or are not similar to the โtrueโ, represented in our case by the IP92 catalog.
Finally, we would like to stress that in statistical pattern recognition, one of the main problems in evaluating the system performances is the optimisation of all the compared systems in order not to give any unfair advantage to one of the systems with respect to the others (just because it is better optimised than the others). For instance, since the background subtraction is crucial to the detection, all algorithms, including SEx, were run on the same background subtracted image.
### 3.3 Segmentation
From the astronomical point of view, segmentation allows to disentangle objects from noisy background. From a mathematical point of view, instead, the segmentation of an image $`F`$ consists in splitting it into disconnected homogeneous (accordingly to a uniformity predicate $`P`$) regions $`\{S_{1,}\mathrm{},S_n\}`$, in such a way that their union is not homogeneous:
$$\underset{i=1}{\overset{n}{}}S_i=F\mathrm{with}S_iS_j=\mathrm{},ij$$
where $`P(S_i)=true`$ $`i`$ and $`P(S_iS_j)=false`$ when $`S_i`$ is adjacent to $`S_j`$. The two regions are adjacent when they share a boundary, i.e. when they are neighbours.
A segmentation problem can be easily transformed into a classification one if classes are defined on pixels and $`P`$ is written in such a way that $`P(S_i)=true`$ if and only if all the pixels of $`S_i`$ belong to the same class. For instance, the segmentation of an astronomical image in background and objects leads to assign each pixel to one of the two classes. Among the various methods discussed in the literature, unsupervised NNโs usually provide better performance than any other NN type on noisy data (Pal & Pal 1993) and have the great advantage of not requiring a definition (or exhaustive examples) of object.
The first step of the segmentation process consists in creating a numerical mask where different values discriminate the background from the object (Fig. 5).
In well sampled images, the attribution of a pixel to either the background or to the object classes depends on both the pixel value and on the properties of its neighbours: for instance, a โbrightโ isolated pixel in a โdarkโ environment is usually just noise. Therefore, in order to classify a pixel, we need to take into account the properties of all the pixels in a $`(n\times n)`$ window centered on it. This approach can be easily extended to the case of multiband images. $`n\times n`$, however, is a too high dimensionality to be effectively handled (in terms of learning and computing time) by any classification algorithm. Therefore, in order to lower the dimensionality, we first use a PCA to identify the $`M`$ (with $`M<<n\times n`$) most significant features. In detail:
i) we first run the $`\left(n\times n\right)`$ window on a sub-image containing representative parts of the image. We used both a $`3\times 3`$ and a $`5\times 5`$ windows.
ii) Then we train the PCA NNโs on these patterns. The result is a projection matrix $`W`$ with dimensionality $`\left(n\times n\right)\times M`$, which allows us to reduce the input feature number from $`\left(n\times n\right)`$ to $`M`$. We considered only the first three components since, accordingly to the PCA, they contain almost $`93\%`$ of the information while the remaining $`7\%`$ is distributed over all the others.
iii) The $`M`$-dimensional projected vector $`WI`$ is the input of a second NN which classifies the pixels in the various classes.
iv) Finally, we merge all classes except the background one in order to reduce the classification problem to the usual โobject/backgroundโ dichotomy.
Much attention has also to be paid to the choice of the type of PCA. After several experiments, we found that - for our specific task which is characterised by a large dynamical range in the luminosities of the objects (or, which is the same, in the pixel values) - PCAโs can be split into two gross groups: PCAโs with linear input-output mapping (hereafter linear PCA NNโs) and PCAโs with non linear input-output mapping (non-linear PCA NNโs) (see section 2.1). Linear PCA NNโs turned out to misclassify faint objects as background. Non-linear PCA NNโs based on a sigmoidal function allowed, instead, the detection of faint sources. This can be better understood from Fig. 6 and 7 which give the distributions of the training points in the simpler case of two dimensional inputs for the two types of PCA NNโs.
Linear PCA NNโs produce distributions with a very dense core (background and faint objects) and only a few points (luminous objects) spread over a wide area. Such a behaviour results from the presence of very luminous objects in the training set which compress the faint ones to the bottom of the scale. This problem can be circumvented by avoiding very luminous objects in the training set, but this would make the whole procedure too much dependent on the choice of the training set.
Non-linear PCA NNโs, instead, produce better sampled distributions and a better contrast between background and faint objects. The sigmoidal function compresses the dynamical range squeezing the very luminous objects into a narrow region (see Fig. 7).
Among all, the best performing NN (Tagliaferri et al. 1998) turned out to be the hierarchical robust PCA NN with learning function $`g^{(t)}=tanh(\alpha x)`$ given in Eq. 5. This NN was also the faster among the other non-linear PCA NNโs.
The principal components matrices are detailed in the tables 1-3 and 4-6 for the $`3\times 3`$ and $`5\times 5`$ cases, respectively. In tables 1โ3, numbers are rounded to the closest integere since they differ from an integer only at the 7-th decimal figure. Not surprisingly, the first component turns out to be the mean in the $`3\times 3`$ case. The other two matrices can be seen as anti-symmetric filters with respect to the centre. The convolution of these filters (see Fig. 8) with the input image gives images where the objects are the regions of high contrast. Similar results are obtained for the $`5\times 5`$ case.
At this stage we have the principal vectors and, for each pixel, we can compute the values of the projection of each pixel in the eigenvector space. The second step of the segmentation process consists in using unsupervised NNโs to classify the pixels into few classes, having as input the reduced input patterns which have been just computed. Supervised NN would require a training set specifying, for each pixel, whether that pixel belongs to an object or to the background. We no longer consider such a possibility, due to the arbitrariness of such a choice at low fluxes, the lack of elegance of the method and the problems which are encountered in the labelling phase. Unsupervised NNโs are therefore necessary. We considered several types of NNโs.
As already mentioned several times, our final goal is to classify the image pixels in just two classes: objects and background, which should correspond to two output neurons. This simple model, however, seldom suffice to reproduce real data in the bidimensional case (but similar results are obtained also for the 3-D or multi-D cases), since any unsupervised algorithm fails to produce spatially well separated clusters and more classes are needed. A trial and error procedure shows that a good choice of classes is $`6`$: fewer classes produce poor classifications while more classes produce noisy ones. In all cases, only one class (containing the lowest luminosity pixels) represents the background, while the other classes represent different regions in the objects images.
We compared hierarchical, hybrid and unsupervised NNโs with $`6`$ output neurons. From theoretical considerations and from preliminary work (Tagliaferri et. al 1998) we decided to consider only the best performing NNโs, id est Neural gas, ML-Neural gas, ML-SOM, and GCS+ML-Neural gas. For a more quantitative and detailed discussion see section 3.6, where the performances of these NNโs are evaluated.
After this stage all pixels are classified in one of six classes. We merge together all classes, with the exception of the background one and reduce the classification to the usual astronomical dichotomy: object or background.
Finally, we create the masks, each one identifying one structure composed by one or more objects. This task is accomplished by a simple algorithm, which, while scanning the image row by row, when it finds one or more adjacent pixels belonging to the object class expands the structure including all equally labelled pixels adjacent to them.
Once objects have been identified we measure a first set of parameters. Namely: the photometric barycenter of the objects computed as:
$$\overline{x}=\frac{_{(x,y)A}xI(x,y)}{flux}\mathrm{and}\overline{y}=\frac{_{(x,y)A}yI(x,y)}{flux}$$
where $`A`$ is the set of pixels assigned to the object in the mask, $`I(x,y)`$ is the intensity of the pixel $`(x,y)`$, and
$$flux=\underset{(x,y)A}{}I(x,y)$$
is the flux of the object integrated over the considered area. The semimajor axis of the object contour defined as:
$$a=\underset{(x,y)A}{\mathrm{max}}(x,y)(\overline{x},\overline{y})=\underset{(x,y)A}{\mathrm{max}}r(x,y)$$
with position angle defined as:
$$\alpha =\mathrm{arctan}\left(\frac{y^{}\overline{y}}{x^{}\overline{x}}\right)$$
where $`(x^{},y^{})`$ is the most distant pixel from the barycenter belonging to the object. The semiminor axis of the faintest isophote is given by:
$$b=\underset{(x,y)A}{\mathrm{max}}\left|\mathrm{sin}\left[\mathrm{arctan}\left(\frac{y\overline{y}}{x\overline{x}}\right)\alpha \right]\right|r(x,y)$$
These parameters are needed in order to disentangle overlapping objects.
### 3.4 Object deblending
Our method recognises multiple objects by the presence of multiple peaks in the light distribution. Search for double peaks is performed along directions at position angles $`\beta _i=\alpha +i\pi /n`$ with $`0in`$. At difference with FOCAS, (Jarvis & Tyson 1981), we sample several position angles because not always objects are aligned along the major axis of their light distribution, as FOCAS implicitly assumes. In our experiments the maximum $`n`$ was set to $`5`$. When a double peak is found, the object is split into two components by cutting it perpendicularly to the line joining the two peaks.
Spurious peaks can also be produced by noise fluctuations, a case which is very common in photographic plates near saturated objects. A good way to minimise such noise effects is, just for deblending purposes, to reduce the dynamical range of the pixels values, by rounding the intensity (or pixel values) in $`N`$ equi-espaced levels.
Multiple (i.e. $`3`$ or more components) objects pose a more complex problem. In the case shown in Fig. 9, the segmentation mask includes three partially overlapping sources. The search for double peaks produces a first split of the mask into two components which separate the third and faintest component into two fragments. Subsequent iterations would usually produce a set of four independent components therefore introducing a spurious detection. In order to solve the problem posed by multiple โnon spuriousโ objects erroneously split, a recomposition loop needs to be run. Most celestial objects - does not matter whether resolved or unresolved - present a light distribution rapidly decreasing outwards from the centre. If an object has been erroneously split into several components, then the adjacent pixels on the corresponding sides of the two masks will have very different values. The implemented algorithm checks each component (starting from the one with the highest average luminosity and proceeding to the fainter ones) against the others. Let us now consider two parts of an erroneously split object. When the edge pixels have luminosity higher than the average luminosity of the faintest component, the two parts are recomposed. This procedure also takes care of all spurious components produced by the haloes of bright objects (an artifact which is a major shortcoming of many packages available in the astronomical community).
### 3.5 Contour regularisation
The last operation before measuring the objects parameters consists in the regularization of the contours since โ due to noise, overlapping images, image defects, etc. โ segmentation produces patterns that are not similar to the original celestial objects that they must represent. For the contour regularisation, we threshold the image at several sigma over the background and we then expand the ellipse describing the objects in order to include the whole area measured in the object detection.
### 3.6 Results on the object detection phase
After the above described steps, it becomes possible to measure and compare the performances of the various NN models. We implemented and compared: Neural Gas (NG3), ML-Neural Gas (MLNG3 or MLNG5), ML-SOM (K5), GCS+ML-Neural Gas (NGCS5). The last digit in the NN name indicating the dimensions of the running window.
Attention was paid in choosing the training set, which needed to be at the same time small but significant. By trial and error, we found that for PCA NNโs and unsupervised NNโs it was enough to choose $`10`$ sub-images, each one $`50\times 50`$ pixels wide and not containing very large objects. As all the experienced users know, the choice of the SEx parameters (minimum area, threshold in units of the background noise, and deblending parameter) is not critical and the default values were choosen (4 pixel area, $`1.5\sigma `$).
Table 7 shows the number of objects detected by the five NNโs and SEx. It has to be stressed that $`2100`$ objects out of the 4819 available in the IP92 reference catalogue are beyond the detection limit of our plate material. SEx detects a larger number of objects but many of them (see Table 7) are spurious. NNโs detect a slightly smaller number of objects but most of them are real. In particular: MNG5 looses, with respect to SEx, only 79 real objects but detects 400 spurious objects less; MNG3 is a little less performing in detecting true objects but is even cleaner of spurious detections.
The upper panel of Fig. 10 shows the number of โTrueโ objects (i.e. objects in the IP92 catalogue). Most of them are fainter than $`m_F21.5`$ mag, id est they are fainter than the plate limit. The lower panel shows instead the number of objects detected by the various NNโs relative to SEx. The curves largely coincide and, in particular, MLNG5 and SEx do not statistically differ in any magnitude bin while MLNG3 slightly differs only in the faintest bin ($`m_F21.5`$).
The class of โMissedโ objects (id est objects which are listed in the reference catalogue but are not in the NNโs or SEx catalogues) needs a detailed discussion. We focus first on brighter objects. They can be divided in:
โ Few โTrueโ objects with a nearby companion which are blended in our image but are resolved in IP92.
โ Parts of isolated single large objects incorrectly split by IP92. A few cases.
โ A few detections aligned in the E-W direction on the two sides of the images of a bright star. They are likely false objects (diffraction spikes detected as individual objects in the IP92 catalog).
โ Objects in IP92 which correspond to empty regions in our images: they can be missing because variable, fast moving, or with an overestimated luminosity in the reference catalog; they can also be missed because spurious in the template catalog.
Therefore, a fair fraction of the โMissedโ objects is truly non existent and the performances of our detection tools are lower bounded at $`m_F<21`$ mag. We wish to stress here that even though there is nothing like a perfect catalogue, the IP92 template is among the best ones ever produced to our knowledge.
The upper panel of Fig. 11 is the same as in Fig. 10. The lower panel shows instead the fraction of โfalseโ objects, id est of the objects detected by the algorithms but not present in the reference catalogue. IP92 were interested to faint objects and masked out the bright ones, therefore their catalogue may exclude a few โTrueโ objects (in particular at $`m_F17`$). We believe that all objects brighter than $`m_F=20`$ mag are really โTrueโ since they are detected both by SEx and NNโs with high significance. For objects brighter than $`m_F=20`$ mag, the NNโs and SEx have similar performances. They differ only at fainter magnitudes. The catalogue with the largest contamination by โFalseโ objects is SEx, followed by MLNG5, MLNG3 and the other NNโs beeing much less contaminated. MLNG5 is quite efficient in detecting โTrueโ objects and has a $`20\%`$ cleaner detection rate in the highly populous bin $`m_F=21.7`$ mag. MLNG3 is less efficient than MLNG5 in detecting โTrueโ objects but it is even cleaner than MLNG5 of false detections.
Let us now consider whether or not the detection efficiency depends on the degree of concentration of the light (stars have steeper central gradients than galaxies). In IP92 objects are classified in two major classes, star & galaxies, and a few minor ones (merged, noise, spike, defects, etc.) that we neglect. The efficiency of the detection is shown in Fig. 12 for three representative detection algorithms: MLNG5, K5, and SEx. At $`m_F<21`$ mag, the detection efficiency is large, close to 1 and independent on the central concentration of the light. Please note that there are no objects in the image having $`m_F<16`$ mag and that in the first bin there are only 4 galaxies. At fainter magnitudes ($`2223`$ mag) detection efficiencies differ as a function of both the algorithm and of the light concentration. In fact, SEx, MLNG5, and to less extent K5, turn out to be more efficient in detecting galaxies rather than stars (in other words: โMissedโ objects are preferentially stars). For SEx, a possible explanation is that a minimal area above the background is required in order for the object to be detected and at $`m_F2223`$ mag and noise fluctuations can affect the isophotal area of unresolved objects bringing it below the assumed threshold (4 pixels). This bias is minimum for the K5 NN. However, this is more likely due to the fact that K5 misses more galaxies than the other algorithms, rather than to the fact that it detects more stars.
In conclusion: MLNG3 and MLNG5 turn out to have high performances in detecting objects: they produce catalogs which are cleaner of false detections at the price of a slightly larger uncompleteness than the SEx catalogues below the plate completness magnitude.
We also want to stress that since the less performing NNโs produce catalogs which are much cleaner of false detections, the selected objects are in large part โtrueโ, and not just noise fluctuations. These NNโs can therefore be very suitable to select candidates for possible followโup detailed studies at magnitudes where many of the objects detected by SEx would be spurious. Deeper catalogs having a large number of spurious source, such as those produced by SEx or other packages are instead preferable if, for instance, they can be cleaned by subsequent processing (for instance by matching the detected objects with other catalogs).
A posteriori, one could argue that performances similar to those of each of the NNโs could be achieved by running SEx with appropriate settings. However, it would be unfair (and methodologically wrong) to make a fine tuning of any of the detection algorithms using an a-posteriori knowledge. It would also make cumbersome the authomatic processing of the images which is the final goal of our procedure.
### 3.7 Feature extraction and selection
In this section we discuss the feature extraction and selection of the features which are useful for the star/galaxy classification. Features are chosen from the literature (Jarvis & Tyson 1981; Miller & Coe 1996; Odewahn et al. 1992 (=O92), Godwin & Peach 1977), and then selected by a sequential forward selection process (Bishop 1995), in order to extract the most performing ones for classification purposes.
The first five features are those defined in the previous section and describing the ellipses circumscribing the objects: the photometric barycenter coordinates ($`\overline{x},\overline{y}`$), the semimajor axis ($`a`$), the semiminor axis ($`b`$). and the position angle ($`\alpha `$). The sixth one is the object area, $`A`$, i.e. the number of pixels forming the object.
The next twelve features have been inspired to the pioneeristic work by O92: the object diameter ($`dia=2a`$), the ellipticity ($`ell=1b/a`$), the average surface brightness ($`SuBr=\frac{1}{A}_{(x,y)A}I(x,y)`$), the central intensity ($`I_0=I(\overline{x},\overline{y})`$), the filling factor ($`f_{fac}=\pi ab/A`$), the area logarithm ($`c_2=\mathrm{log}\left(A\right)`$), the harmonic radius ($`r_1`$). The latter beeing defined as:
$$r_1=\frac{1}{flux}\underset{(x,y)A}{}\frac{I(x,y)}{r(x,y)}$$
and five gradients $`G_{14}`$, $`G_{13}`$, $`G_{12}`$, $`G_{23}`$ and $`G_{34}`$ defined as:
$$G_{ij}=\frac{T_jT_i}{r_ir_j}$$
where $`T_i`$ is the average surface brightness within an ellipse, with position angle $`\alpha `$, semimajor axis $`r_i=ia/4`$, $`i=1,\mathrm{}4`$. and ellipticity $`ell`$.
Two more features are added following Miller & Coe (1996): the ratios $`T_r=SuBr/I_0`$ and $`T_{cA}=I_0/\sqrt{A}`$.
Finally, five FOCAS features (Jarvis & Tyson 1981) have been included: the second ($`C_2`$) and the fourth ($`C_4`$) total moments defined as:
$$C_2=\frac{M_{20}+M_{02}}{M_{00}}\mathrm{and}C_4=\frac{M_{40}+2M_{22}+M_{04}}{M_{00}}$$
where $`M_{ij}`$ are the object central momenta computed as:
$$M_{ij}=\underset{(x,y)A}{}\left(x\overline{x}\right)^i\left(y\overline{y}\right)^jI(x,y),$$
the average ellipticity:
$$E=\frac{\sqrt{\left(M_{20}M_{02}\right)^2+M_{11}^2}}{M_{02}+M_{20}},$$
the central intensity averaged in a $`3\times 3`$ $`area`$ and, finally, the โKronโ radius defined as:
$$r_{Kron}=\frac{1}{flux}\underset{(x,y)A}{}I(x,y)r(x,y)$$
For each object we therefore measure $`25`$ features, where the first $`6`$ are reported only to easy the graphical representation of the objects and have a low discriminating power. The complete set of the extracted features is given in Table 8.
Our list of features includes therefore most of those usually used in the astronomical literature for the star/galaxy classification.
Are all these features truly needed? And, if this is not and a smaller subset contains all the needed information, what are the most useful ones? We tried to answer these questions by evaluating the classification performance of each set of features through the a-priori knowledge of the true classification of each object, as it is listed in a much deeper and higher quality reference catalog.
Most of the defined features are not independent. The presence of redundant features decreases the classification performances since any algorithm would try to minimise the error with respect to features which are not particularly relevant for the task. Furthermore, by introducing useless features the computational speed would be lowered.
The feature selection phase was realised through the sequential backward elimination strategy (Bishop 1995), which works as follows: let us suppose to have $`M`$ features in one set and to run the classification phase with this set. Then, we build $`M`$ different sets with $`M1`$ features in each one and then we run the classification phase for each set, keeping the set which attains the best classification. This procedure allows us to eliminate the less significant feature. Then, we repeat $`M1`$ times the procedure eliminating one feature at each step. In order to further reduce the computation time we do not use the validation set and the classification error is evaluated directly on the test set. It has to be stressed that this procedure is common in the statistical pattern recognition literature where, very often, for this task are also introduced simplified models. This however could be avoided in our case due to the speed and good performances of our NNโs
Unsupervised NNโs were not successful in this task, because the input data feature space is not separated into two not overlapping classes (or, in simpler terms, the images and therefore the parameters of stars and galaxies fainter than the completeness limit of the image are quite similar), and they reach a performance much lower than supervised NNโs.
Supervised learning NNโs give far better results. We used a MLP with one hidden layer of $`19`$ neurons and only one output, assuming value $`0`$ for star and value $`1`$ for galaxy. After the training, we calculate the NN output as $`1`$ if it is greater than $`0.5`$ and $`0`$ otherwise for each pattern of the test set. The experiments produce a series of catalogues, one for each set of features.
Fig. 13 shows the classification performances as a function of the adopted features. After the first step, the classification error remains almost constant up to $`n=4`$, id est up to the point where features which are important for the classification are removed.
A high performance can be reached using just 6 features. With a lower number of features the classification worsen, whereas a larger number of features is unjustified, because it does not increase the performances of the system. The best performing set of features consists of features 11, 12, 14, 19, 21, 25 of table 8. They are two radii, two gradients, the second total moment and a ratio which involves measures of intensity and area.
### 3.8 Star/Galaxy classification
Let us discuss now how the star/galaxy classification takes place. The first step is accomplished by โteachingโ the MLP NN using the selected best features. In this case we divided the data set into three independent data sets: training, validation and test sets. The learning optimization is performed using the training set while the early stopping technique (Bishop 1995) is used on the validation set to stop the learning to avoid overfitting. Finally, we run the MLP NN on the test set.
As comparison classifier, we adopt SEx, which is based on a MLP NN. As features useful for the classification, SEx uses eight isophotal areas and the peak intensity plus a parameter, the FWHM of stars. Since the SEx NN training was already realised by Bertin & Arnouts (1996) on $`10^6`$ simulated images of stars and galaxies, we limit ourselves to tune SEx in order to obtain the best performances on the validation set. Both SEx and our system use NNโs for the classification, but they follow two different, alternative approaches: SEx uses a very large training set of simulated stars and galaxies, our system uses noisy, real data. Furthermore, while the features of SEx are fixed by the authors, and the NNโs output is a number $`x`$, $`0<x<1`$; our system selects the best performing ones and its output is an integer: 0 or 1 (id est star or galaxy). Therefore, we use the validation set for choosing the threshold which maximises the number of correct classifications by SEx (see Fig. 14).
The experimental results are shown in Fig. 15 where the errors are plotted as a function of the magnitude. At all magnitudes NExt misclassify less objects than SEx. Out of 460 objects, SEx makes 41 incorrect classifications, while NExt just 28.
In order to check the that our feature selection is optimal, we also compared our classification with those obtained using our MLP NNโs with others feature sets, selected as shown in Table 9. The total number of misclassified objects in the test set of 460 elements were: O-F, 43 errors; O-L, 30 errors; O-S, 35 errors; GP1, 48 errors; GP2, 49 errors. Fig. 16 shows the classification performances of the considered feature sets as a function of the magnitude of the objects. Results for stars are presented as solid line, while for galaxies we used dotted lines. The perfomances of NExt are presented in the top-left panel: galaxies are correctly classified as long as they are detected, whereas the correctness of the classification of stars drops to 0 at $`m_F=21`$. Fainter stars are pratically absent in the IP92 catalog, thus explaining why the stars point stop at brighter magnitudes than galaxies. O92 selected a 9 features set (O-F) for the star/galaxy classification. Their set (centralโleft panel) is slightly less performing for bright ($`m_F=17`$) galaxies and for faint stars ($`m_F=19`$) than the set of features selected by us (upperโleft panel). They select also a smaller (four) set of features (O-F) quite useful to classify large objects. The classification performances of this set, when applied to our images, turn out to be better than the larger feature dataset: in fact, bright galaxies are not misclassified (see the bottom left panel). Even with respect to our dataset O-F performs well: their set is sligthly better in classifying bright galaxies, at the price of a achieving lower performances on faint stars. The further set of features by O92 (O-S) was aimed to the accurate detection of faint sources and performs similarly to their full set: it misclassifies bright galaxies and faint stars. The performances of the traditional classifiers, $`magvsarea`$ (GP1) and $`magvsbrightness`$ (GP2), are presented in the central and low right panels. With just two features, all the faint objects are classified as galaxies, and due to the absence of stars in our reference catalog, the classification performances are $`100\%`$. However, this is not a real classification. At bright magnitudes, the classification of the traditional classified dataset are as large as, or sligthly lower, than the NExt dataset.
## 4 Summary and conclusions
In this paper we discuss a novel approach to the problem of detection and classification of objects on WF images. In Section 2 we shortly review the theory of some type of NNโs which are not familiar to the astronomical community. Based on these considerations, we implemented a Neural Network based procedure (NExt) capable to perform the following tasks: i) to detect objects against a noisy background; ii) to deblend partially overlapping objects; iii) to separate stars from galaxies. This is achieved by a combination of three different NNโs each performing a specific task. First we run a non linear PCA NN to reduce the dimensionality of the input space via a mapping from pixels intensities to a subspace individuated through principal component analysis. For the second step we implemented a hierarchical unsupervised NN to segmentate the image and, finally after a deblending and reconstruction loop we implemented a supervised MLP to separate stars from galaxies.
In order to identify the best performing NNโs we implemented and tested in homogeneous conditions several different models. NExt offers several methodological and practical advantages with respect to other packages: i) it requires only the simplest a priori definition of what an โobjectโ is; ii) it uses unsupervised algorithms for all those tasks where both theory and extensive testing show that there is no loss in accuracy with respect to supervised methods. Supervised methods are in fact used only to perform star/galaxy separation since, at magnitudes fainter than the completeness limit, stars are usually almost indistinguishable from galaxies and the parameters characterizing the two classes do not lay in disconnected subspaces. iii) Instead of using an arbitrarily defined and often specifically tailored set of features for the classification task NExt, after measuring a large set of geometric and photometric parameters, uses a sequential backward elimination strategy (Bishop 1995) to select only the most significant ones. The optimal selection of the features was checked against the performances of other classificators (see Sect. 3.8).
In order to evaluate the performances of NExt, we tested it against the best performing package known to the authors (id est SEx) using a DPOSS field centered on the North Galactic Pole. We want also to stress here that - in order to have an objective test and at difference of what is currently done in literature - NExt was checked not against the performances of an arbitrarily choosen observer but rather against a much deeper catalogue of objects obtained from better quality material.
The comparison of NExt performances against those of SEx show that in the detection phase, NExt is at least as effective as SEx in detecting โtrueโ objects but much cleaner of spurious detections. For what classification is concerned, NExt NN performs better than the SEX NN: 28 errors for NExt against 41 for SEx on a total of 460 objects, most of the errors referring to objects fainter than the plate detection limit.
Other attempts, besides those described in the previous sections, to use NN for similar tasks have been discussed in the literature. Balzell & Peng (1998), used the same North Galactic Pole field (but extracted from POSS-I plates) used in this work. They tested their star/galaxy classification NN on objects which are both too few (60 galaxies and 27 stars) and too bright (a random check of their objects shows that most of the galaxies extend well over than 20 pixels) to be of real interest. It needs also to be stressed that, due to their preprocessing strategy, their NNโs are forced to perform cluster analysis on a huge multidimensional imput space with scarsely populated samples.
Naim (1997) follows instead a strategy which is similar to ours and makes use of a fairly large dataset extracted from POSS-I material. He, however, trained the networks to achieve the same performances of an experienced human observer while, as already mentioned, NExt is checked against a catalogue of โTrueโ objects. Even though his target is the classification of objects fainter and larger than those we are dealing with, he tested the algorithm in a much more crowded and difficult region of the sky near the Galactic plane.
O92 makes use of a traditional MLP and succeeded in demonstrating that AI methods can reproduce the star/galaxy classification obtained with traditional diagnostic diagrams by trained astronomers. Their aim, however, was less ambitious than that of โperforming the correct star/galaxy classificationโ which is instead the final goal of NExt.
This paper is a first step toward the application of Artificial Intelligence methods to astronomy. Foreseen improvements of our approach are the use of ICA (Independent Component Analysis) NNโs instead of PCA NNโs and the adoption of Bayesian learning techniques to improve the classification performences of MLPโs. These developments and the application of NExt to other wide field astronomical data sets obtained at large format CCD detectors will be discussed in forthcoming papers.
Acknowledgements The authors wish to thank Chris Pritchet for providing them with a digital version of the IP92 catalogue. We also acknoledge the Canadian Astronomy Data Center for providing us with POSS-II material. This work was partly sponsored by the special grant MURST COFIN 1998, n.9802914427.
|
warning/0006/hep-th0006115.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The idea that observable Universe represents the brane where 4d gravity is trapped embedded in higher dimensional bulk space attracted enormous attention. The study of cosmological aspects of such brane-worlds (and refs. therein) indicates towards the possibility to construct the inflationary brane Universe . Some observational manifestations of bulk matter fields may occur on the brane level . As brane-world Universe is naturally realized in more than four dimensional bulk space it could be related with recent studies of AdS/CFT correspondence . One attempt could be in implementing of RS warped compactification within the context of RG flow in AdS/CFT set-up. It may be done in the simplest form in the way suggested in refs. via consideration of quantum CFT living on the brane.
Indeed, this scenario significally varies from original approach where brane tension (brane cosmological constant) is free parameter of theory permitting actually the existence of brane solution by its fine-tuning. In scenario the action is defined from the beginning and surface terms are added in closed manner. The role of surface terms is to make the variational procedure to be well-defined and to cancell the leading divergences of the action around AdS bulk space. Hence, brane tension is not free parameter anymore. However, its role is taken by quantum effects. Indeed, quantum effects of brane CFT induce the conformal anomaly and anomaly induced effective action. This 4d anomaly induced (brane) effective action should be added to the complete effective action of five-dimensional theory. In fact, it gives explicit contribution to brane tension permitting to have the consistent curved brane solutions. As the result, the possibility of quantum creation of de Sitter or Anti-de Sitter brane living in 5d AdS Universe has been proved in refs.. The simplest choice for brane CFT is maximally supersymmetric (SUSY) Yang-Mills theory. Thus, conformal anomaly of brane CFT induces the brane effective tension which is responsible for de Sitter (or AdS) geometry of brane. Note that anomaly induced effective action may be considered as kind of gravitational Casimir effect (for a recent introduction to Casimir effect, see ).
Developing further the study of warped compactifications with curved boundary (inflationary brane) within AdS/CFT correspondence the natural question is about the role of quantum bulk effects in such scenario. In other words, quantum effects of five-dimensional bulk gravity (i.e. bulk gravitational Casimir effect) should be taken into account. When the boundary of AdS space is flat it can be done in the analogy with the usual calculation of quantum effective action in Kaluza-Klein multi-dimensional gravity (for a review and complete list of references, see ). However, for AdS bulk space with curved boundary such quantum gravity calculation is much more involved. This will be done elsewhere.
In the present paper, in order to estimate (at least, qualitatively) the role of bulk quantum effects to the scenario of refs. we consider the contribution of quantum bulk matter (on the example of scalar) to complete five-dimensional effective action. Having the structure of bulk effective action for conformal matter we discuss brane-world cosmology where bulk and boundary quantum effects are taken into account. The effective bulk-brane equations of motion are derived and their solutions are analyzed. It is shown that due to such effects the quantum creation of de Sitter or Anti-de Sitter brane living in 5d AdS Universe is possible, where bulk quantum effects deform the shape of constant curvature brane (our observable Universe). Note that such Universe occurs when only brane quantum effects are included. Hence, bulk quantum effects modify the geometrical configuration which is recovered in their absence. It is interesting that above quantum creation is possible only due to bulk matter quantum effects,i.e. when no CFT lives on the brane.
## 2 Gravitational Casimir effect for bulk scalar in AdS space
Let us start from the following lagrangian (Euclidean sector) for a conformally invariant massless scalar field with scalar-gravitational coupling
$$=\sqrt{g}\chi (\mathrm{}+\xi R)\chi $$
(1)
where $`\xi =\frac{๐2}{4(๐1)}`$. Having in mind the applications to Randall-Sundrum scenario one takes $`๐`$ to be odd. First of all one considers the warped metric of the form typical for warped compactification :
$$ds^2=dy^2+\mathrm{e}^{2A(y)}d\mathrm{\Omega }_{๐1}^2=\mathrm{e}^{2A(z)}\left[dz^2+d\mathrm{\Omega }_{๐1}^2\right],$$
(2)
where $`d\mathrm{\Omega }_{๐1}^2`$ corresponds to a $`๐`$-1-dimensional constant curvature symmetric space M<sub>D-1</sub>, namely R<sub>D-1</sub>, S<sub>D-1</sub> and H<sub>D-1</sub>, the Euclidean space, the sphere and the hyperbolic space respectively. Furthermore, the warp factor is given by
$$\mathrm{e}^{A(z)}=\frac{l}{z},$$
(3)
$`l`$ being related to the cosmological constant. We may put $`l=1`$ and later by dimensional analysis recover it. Furthermore, in the following, let us consider $`๐=5`$. It may be convenient to make the conformal transformation for the metric
$$(ds^2)^{}=z^2ds^2$$
(4)
and for the scalar field $`\chi ^{}=z^{3/2}\chi `$. Then, for rescaled scalar $`\chi ^{}`$, one gets the Lagrangian
$$=\chi ^{}(\mathrm{}^{}+\xi R_{(4)})\chi ^{},$$
(5)
where $`R_{(4)}`$ is the constant scalar curvature of M<sub>4</sub> and $`\mathrm{}^{}`$ is the Laplace operator in the product space $`\mathrm{R}\times \mathrm{M}_4`$, whose domain is subject to suitable boundary conditions at $`z=l`$ and $`z=L`$, induced by the orbifold nature of the space-time we have started with.
Now, one should calculate the one-loop effective potential, i.e. the effective action divided by the whole volume
$$V=\frac{1}{2L\mathrm{Vol}(\mathrm{M}_4)}\mathrm{log}det\left(L_5/\mu ^2\right),$$
(6)
where
$$L_5=_z^2\mathrm{}_{(4)}+\xi R_{(4)}=L_1+L_4,$$
(7)
on $`\mathrm{R}\times \mathrm{M}_4`$ limited by two branes subject to boundary conditions. At the end, if $`L`$ is the branes separation, we will take the limit $`L`$ goes to infinity.
Since, one is dealing with a product space, the heat-kernel for $`L_5`$ is given by
$$K_t(L_5)=K_t(L_1)K_t(L_4),$$
(8)
where
$$K_t(L_1)=\underset{n}{}\mathrm{e}^{t\lambda _n^2}$$
(9)
is the heat-kernel related to the one-dimensional operator $`L_1=_z^2`$ whose domain contains the conformally transformed orbifold boundary conditions, $`\lambda _n`$ the associated eigenvalues and
$$K_t(L_4)=\underset{\alpha }{}\mathrm{e}^{t\lambda _\alpha ^2}$$
(10)
is the heat-kernel of the Laplace-like operator $`L_4`$ on M<sub>4</sub>, $`\lambda _\alpha =\mu _\alpha +\xi R_{(4)}`$, $`\mu _\alpha `$ being the eigenvalues of $`\mathrm{}_{(4)}`$.
If we make use of the zeta-function regularization, one needs the analytical continuation of the zeta-function
$$\zeta (s|L_5)=\underset{\alpha }{}\underset{n=1}{\overset{\mathrm{}}{}}(\lambda _n+\lambda _\alpha ^2)^s.$$
(11)
Generally speaking, we do not know explicitly the spectrum of $`L_1`$. However, it is known the short $`t`$ asymptotics, which is given by
$$K_t(L_1)=\underset{r=0}{\overset{\mathrm{}}{}}K_r(L_1)t^{\frac{r1}{2}}=\frac{L}{2\sqrt{\pi t}}+K_1(L_1)+O(t^{1/2})+O\left(\frac{1}{L}\right)+O(\mathrm{e}^{\frac{L^2}{t}}).$$
(12)
Here $`L`$ is the brane separation, the leading term is the Weyl term and the next term is the first non-trivial boundary term, which, for dimensional reasons, is a numerical constant. Since there is no potential term, the other boundary terms are, again for dimensional reasons, of order $`O(1/L)`$. Thus, if $`L`$ goes to infinity, the above asymptotics becomes almost exact. As a result, we have
$$\zeta (s|L_5)=\frac{L\mathrm{\Gamma }\left(s\frac{1}{2}\right)}{2\mathrm{\Gamma }(s)}\zeta \left(s\frac{1}{2}|L_4\right)+K_1(L_1)\zeta (s|L_4)+O\left(\frac{1}{L}\right),$$
(13)
where $`\zeta (s|L_4)`$ is the zeta-function associated with the Laplace-like operator on M<sub>4</sub>.
It should be noted that
$$\zeta (0|L_5)=K_1(L_1)\zeta (0|L_4).$$
(14)
Thus, there exists also a non trivial contribution coming from the Jacobian related to the conformal transformation we have performed. We are neglecting for the moment this contribution as we argue later on that it is negligible.
As a consequence, in the large $`L`$ limit, the effective potential reads
$`V`$ $`=`$ $`{\displaystyle \frac{1}{2L\mathrm{Vol}(\mathrm{M}_4)}}[\zeta ^{}(0|L_5)+\mathrm{ln}\mu ^2\zeta (0|L_5)]`$ (15)
$`=`$ $`{\displaystyle \frac{\sqrt{\pi }}{2\mathrm{V}\mathrm{o}\mathrm{l}(\mathrm{M}_4)}}\zeta \left({\displaystyle \frac{1}{2}}|L_4\right)+O\left({\displaystyle \frac{1}{L}}\right),`$
In this limit, the effective potential, reduces to the Casimir energy (vacuum energy) related to the Laplace-like operator on M<sub>4</sub>. This Casimir energy has been calculated in several places (see for example, and references therein). We recall the corresponding results.
First, in the case of flat brane R<sub>4</sub>, the Casimir energy goes like $`(O(1/L^4)`$, thus it is negligible.
For the spherical brane S<sub>4</sub> (with radius $``$), the starting point is
$$\zeta (s|L_4)=g(s)^{2s}$$
(16)
where
$$g(s)=\frac{1}{6}\underset{l=1}{\overset{\mathrm{}}{}}(l+1)(l+2)(2l+3)\left(l^2+3l+\frac{9}{4}\right)^s,$$
(17)
the analytical continuation can be easily done, due to conformal coupling in 5 dimensions and the result is
$$\zeta (s|L_4)=\frac{^{2s}}{3}\left[\zeta _H(2s3,\frac{3}{2})\frac{1}{4}\zeta _H(2s1,\frac{3}{2})\right],$$
(18)
where $`\zeta _H(s,a)`$ is the Hurwitz zeta-function. Thus,
$$\zeta \left(\frac{1}{2}|L_4\right)=\frac{1}{3}\left[\zeta _H(4,\frac{3}{2})\frac{1}{4}\zeta _H(1,\frac{3}{2})\right].$$
(19)
Making use of
$$\zeta _H(m,a)=\frac{B_{m+1}(a)}{m+1},$$
(20)
where $`B_n(x)`$ is a Bernoulli polynomial, one gets $`\zeta (\frac{1}{2}|L_4)=0`$. In this case $`V=0`$. This is also consistent with the result reported in ref. . Note that taking non-conformal coupling constant in the initial Lagrangian changes qualitatively this result, then potential will not be zero anymore. It could be also non-zero for another matter fields.
In the hyperbolic brane H<sub>4</sub>, one has
$$\frac{\zeta (s|L_4)}{\mathrm{Vol}(\mathrm{H}_4)}=\frac{^{2s4}}{4\pi ^2}_0^{\mathrm{}}\lambda ^{2s+1}\frac{(\lambda ^2+\frac{1}{4})}{\mathrm{e}^{2\pi \lambda }+1}๐\lambda .$$
(21)
Thus, the effective potential reads
$$V=\frac{1}{8\pi ^{3/2}^5}_0^{\mathrm{}}\frac{(\lambda ^2+\frac{1}{4})}{\mathrm{e}^{2\pi \lambda }+1}๐\lambda .$$
(22)
The integral can be easily evaluated. One has
$$V=\frac{1}{64\pi ^{5/2}^5}\left[\mathrm{ln}2+\frac{3}{4\pi ^2}\zeta _R(3)\right].$$
(23)
Note that back conformal transformation of this potential should be done, to recover the potential in the original AdS space. This transformation introduces the factor $`\mathrm{e}^{5A}`$ in the above expression.
With regard to the Jacobian factor, one may introduce the interpolating
$$(ds_q^2)=z^{2q}ds^2$$
(24)
with $`q`$ a real parameter such that if $`q=0`$, then $`(ds_0^2)=ds^2`$ and if $`q=1`$, then $`(ds_1^2)=(ds^2)^{}`$. Then (see, for example, )
$$\mathrm{ln}J(g,g^{})=_0^1eq_l^L๐zz^{5(q1)}\sqrt{g_4}d^4x\zeta (0|L_5(q))(z,x),$$
(25)
with
$$L_5(q)=\mathrm{}_{(5)}(q)+\xi R_{(5)}(q)$$
(26)
the conformal scalar operator in the interpolating metric and $`\zeta (0|L_5(q))(z,x)=\frac{K_5}{(4\pi )^{5/2}}`$ is the local zeta-function. In a 5-dimensional manifold without boundary, $`K_5=0`$. In our case, however, we have orbifold boundary conditions and the coefficient $`K_5`$ is not vanishing. It has been recently computed for Robin boundary condition in . Unfortunately the expression given in the above reference is difficult to use in our case.
We may assume that $`K_5(q)`$ is a linear combination of terms
$$\underset{l}{}a_lq^l+b_l(1q)^l.$$
(27)
Then, we have
$$\mathrm{ln}J(g,g^{})=\underset{l}{}_0^1๐q_l^L๐zz^{5(q1)}(a_lq^l+b_l(1q)^l).$$
(28)
This integral gives contributions proportional to $`li(L)`$, where $`li(z)`$ is the logarithmic integral. For $`L`$ very large, the logarithmic integral diverges as $`O(\frac{L}{\mathrm{ln}L})`$. Then, it gives a contribution of order $`O(\frac{1}{\mathrm{ln}L})`$ to the effective potential.
Thus, we presented the explicit example of evaluation of gravitational Casimir effect (effective potential) for conformal bulk scalar on five-dimensional AdS space with 4-dimensional sphere or hyperboloid as a boundary. This calculation gives us an idea about the structure of effective action due to bulk matter quantum effects.
## 3 Quantum creation of de Sitter (Anti-de Sitter) brane-world Universe
We consider the spacetime whose boundary is four-dimensional sphere S<sub>4</sub>, which can be identified with a D3-brane or four-dimensional hyperboloid H<sub>4</sub>. The bulk part is given by 5 dimensional Euclidean Anti-de Sitter space $`\mathrm{AdS}_5`$
$$ds_{\mathrm{AdS}_5}^2=dy^2+\mathrm{sinh}^2\frac{y}{l}d\mathrm{\Omega }_4^2.$$
(29)
Here $`d\mathrm{\Omega }_4^2`$ is given by the metric of S<sub>4</sub> or H<sub>4</sub> with unit radius. One also assumes the boundary (brane) lies at $`y=y_0`$ and the bulk space is given by gluing two regions given by $`0y<y_0`$ (see for more details.)
We start with the action $`S`$ which is the sum of the Einstein-Hilbert action $`S_{\mathrm{EH}}`$, the Gibbons-Hawking surface term $`S_{\mathrm{GH}}`$ , the surface counter term $`S_1`$ and the trace anomaly induced action $`W`$<sup>4</sup><sup>4</sup>4For the introduction to anomaly induced effective action in curved space-time (with torsion), see section 5.5 in .:
$`S`$ $`=`$ $`S_{\mathrm{EH}}+S_{\mathrm{GH}}+2S_1+W`$ (30)
$`S_{\mathrm{EH}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G}}{\displaystyle d^5x\sqrt{g_{(5)}}\left(R_{(5)}+\frac{12}{l^2}\right)}`$ (31)
$`S_{\mathrm{GH}}`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G}}{\displaystyle d^4x\sqrt{g_{(4)}}_\mu n^\mu }`$ (32)
$`S_1`$ $`=`$ $`{\displaystyle \frac{3}{8\pi G}}{\displaystyle d^4x\sqrt{g_{(4)}}}`$ (33)
$`W`$ $`=`$ $`b{\displaystyle d^4x\sqrt{\stackrel{~}{g}}\stackrel{~}{F}A}`$ (34)
$`+b^{}{\displaystyle }d^4x\sqrt{\stackrel{~}{g}}\{A[2\stackrel{~}{\mathrm{}}^2+\stackrel{~}{R}_{\mu \nu }\stackrel{~}{}_\mu \stackrel{~}{}_\nu {\displaystyle \frac{4}{3}}\stackrel{~}{R}\stackrel{~}{\mathrm{}}^2+{\displaystyle \frac{2}{3}}(\stackrel{~}{}^\mu \stackrel{~}{R})\stackrel{~}{}_\mu ]A`$
$`+(\stackrel{~}{G}{\displaystyle \frac{2}{3}}\stackrel{~}{\mathrm{}}\stackrel{~}{R})A\}`$
$`{\displaystyle \frac{1}{12}}\left\{b^{\prime \prime }+{\displaystyle \frac{2}{3}}(b+b^{})\right\}{\displaystyle d^4x\sqrt{\stackrel{~}{g}}\left[\stackrel{~}{R}6\stackrel{~}{\mathrm{}}A6(\stackrel{~}{}_\mu A)(\stackrel{~}{}^\mu A)\right]^2}.`$
Here the quantities in the 5 dimensional bulk spacetime are specified by the suffices <sub>(5)</sub> and those in the boundary 4 dimensional spacetime are by <sub>(4)</sub>. The factor $`2`$ in front of $`S_1`$ in (30) is coming from that we have two bulk regions which are connected with each other by the brane. In (32), $`n^\mu `$ is the unit vector normal to the boundary. In (34), one chooses the 4 dimensional boundary metric as
$$g_{(4)}^{}{}_{\mu \nu }{}^{}=\mathrm{e}^{2A}\stackrel{~}{g}_{\mu \nu }$$
(35)
and we specify the quantities with $`\stackrel{~}{g}_{\mu \nu }`$ by using $`\stackrel{~}{}`$. $`G`$ ($`\stackrel{~}{G}`$) and $`F`$ ($`\stackrel{~}{F}`$) are the Gauss-Bonnet invariant and the square of the Weyl tensor <sup>5</sup><sup>5</sup>5We use the following curvature conventions: $`R`$ $`=`$ $`g^{\mu \nu }R_{\mu \nu }`$ $`R_{\mu \nu }`$ $`=`$ $`R_{\mu \lambda \nu }^\lambda `$ $`R_{\mu \rho \nu }^\lambda `$ $`=`$ $`\mathrm{\Gamma }_{\mu \rho ,\nu }^\lambda +\mathrm{\Gamma }_{\mu \nu ,\rho }^\lambda \mathrm{\Gamma }_{\mu \rho }^\eta \mathrm{\Gamma }_{\nu \eta }^\lambda +\mathrm{\Gamma }_{\mu \nu }^\eta \mathrm{\Gamma }_{\rho \eta }^\lambda `$ $`\mathrm{\Gamma }_{\mu \lambda }^\eta `$ $`=`$ $`{\displaystyle \frac{1}{2}}g^{\eta \nu }\left(g_{\mu \nu ,\lambda }+g_{\lambda \nu ,\mu }g_{\mu \lambda ,\nu }\right).`$
$`G`$ $`=`$ $`R^24R_{ij}R^{ij}+R_{ijkl}R^{ijkl}`$
$`F`$ $`=`$ $`{\displaystyle \frac{1}{3}}R^22R_{ij}R^{ij}+R_{ijkl}R^{ijkl},`$ (36)
In the effective action (34), with $`N`$ scalar, $`N_{1/2}`$ spinor, $`N_1`$ vector fields, $`N_2`$ ($`=0`$ or $`1`$) gravitons and $`N_{\mathrm{HD}}`$ higher derivative conformal scalars, $`b`$, $`b^{}`$ and $`b^{\prime \prime }`$ are
$`b`$ $`=`$ $`{\displaystyle \frac{N+6N_{1/2}+12N_1+611N_28N_{\mathrm{HD}}}{120(4\pi )^2}}`$
$`b^{}`$ $`=`$ $`{\displaystyle \frac{N+11N_{1/2}+62N_1+1411N_228N_{\mathrm{HD}}}{360(4\pi )^2}},`$
$`b^{\prime \prime }`$ $`=`$ $`0.`$ (37)
As usually, $`b^{\prime \prime }`$ may be changed by the finite renormalization of local counterterm in gravitational effective action. As we shall see later, the term proportional to $`\left\{b^{\prime \prime }+\frac{2}{3}(b+b^{})\right\}`$ in (34), and therefore $`b^{\prime \prime }`$, does not contribute to the equations of motion. For $`๐ฉ=4`$ $`SU(N)`$ super Yang-Mills theory $`b=b^{}=\frac{N^21}{4(4\pi )^2}`$. As one can see until this point the discussion repeats the one presented in ref. where more detail may be found. It is interesting to note that the contribution from brane quantum gravity may be taken into account via the correspondent coefficient in above equation.
We should also note that $`W`$ in (34) is defined up to conformally invariant functional, which cannot be determined from only the conformal anomaly. The conformally flat space is a pleasant exclusion where anomaly induced effective action is defined uniquely. However, one can argue that such conformally invariant functional gives next to leading contribution as mass parameter of regularization may be adjusted to be arbitrary small (or large).
The metric of $`\mathrm{S}_4`$ with the unit radius is given by
$$d\mathrm{\Omega }_4^2=d\chi ^2+\mathrm{sin}^2\chi d\mathrm{\Omega }_3^2.$$
(38)
Here $`d\mathrm{\Omega }_3^2`$ is the metric of 3 dimensional unit sphere. If we change the coordinate $`\chi `$ to $`\sigma `$ by
$$\mathrm{sin}\chi =\pm \frac{1}{\mathrm{cosh}\sigma },$$
(39)
one obtains
$$d\mathrm{\Omega }_4^2=\frac{1}{\mathrm{cosh}^2\sigma }\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right).$$
(40)
On the other hand, the metric of the 4 dimensional flat Euclidean space is
$$ds_{4\mathrm{E}}^2=d\rho ^2+\rho ^2d\mathrm{\Omega }_3^2.$$
(41)
Then by changing the coordinate as
$$\rho =\mathrm{e}^\sigma ,$$
(42)
one gets
$$ds_{4\mathrm{E}}^2=\mathrm{e}^{2\sigma }\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right).$$
(43)
For the 4 dimensional hyperboloid with the unit radius, the metric is
$$ds_{\mathrm{H4}}^2=d\chi ^2+\mathrm{sinh}^2\chi d\mathrm{\Omega }_3^2.$$
(44)
Changing the coordinate $`\chi `$ to $`\sigma `$
$$\mathrm{sinh}\chi =\frac{1}{\mathrm{sinh}\sigma },$$
(45)
one finds
$$ds_{\mathrm{H4}}^2=\frac{1}{\mathrm{sinh}^2\sigma }\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right).$$
(46)
Motivated by (29), (40), (43) and (46), one assumes the metric of 5 dimensional space time as follows:
$$ds^2=dy^2+\mathrm{e}^{2A(y,\sigma )}\stackrel{~}{g}_{\mu \nu }dx^\mu dx^\nu ,\stackrel{~}{g}_{\mu \nu }dx^\mu dx^\nu l^2\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right)$$
(47)
and we identify $`A`$ and $`\stackrel{~}{g}`$ in (47) with those in (35). Then one finds $`\stackrel{~}{F}=\stackrel{~}{G}=0`$, $`\stackrel{~}{R}=\frac{6}{l^2}`$ etc. Due to Eq. (47), the actions in (31), (32), (33), and (34) have the following forms:
$`S_{\mathrm{EH}}`$ $`=`$ $`{\displaystyle \frac{l^4V_3}{16\pi G}}{\displaystyle }dyd\sigma \{(8_y^2A20(_yA)^2)\mathrm{e}^{4A}`$ (48)
$`+(6_\sigma ^2A6(_\sigma A)^2+6)\mathrm{e}^{2A}+{\displaystyle \frac{12}{l^2}}\mathrm{e}^{4A}\}`$
$`S_{\mathrm{GH}}`$ $`=`$ $`{\displaystyle \frac{3l^4V_3}{8\pi G}}{\displaystyle ๐\sigma \mathrm{e}^{4A}_yA}`$ (49)
$`S_1`$ $`=`$ $`{\displaystyle \frac{3l^3V_3}{8\pi G}}{\displaystyle ๐\sigma \mathrm{e}^{4A}}`$ (50)
$`W`$ $`=`$ $`V_3{\displaystyle }d\sigma [b^{}A(2_\sigma ^4A8_\sigma ^2A)`$ (51)
$`2(b+b^{})(1_\sigma ^2A(_\sigma A)^2)^2].`$
Here $`V_3=๐\mathrm{\Omega }_3`$ is the volume or area of the unit 3 sphere.
As it follows from the discussion in the previous section there is also gravitational Casimir contribution due to bulk quantum fields. As one sees on the example of bulk scalar it has typically the following form $`S_{\mathrm{Csmr}}`$
$$S_{\mathrm{Csmr}}=\frac{cV_3}{^5}๐y๐\sigma \mathrm{e}^A$$
(52)
Note that role of (effective) radius of 4d constant curvature space (after back conformal transformation as in previous section) is played by $`\mathrm{e}^A`$. Here $`c`$ is some coefficient whose value and sign depend on the type of bulk field (scalar, spinor, vector, graviton, โฆ) and on parameters of bulk theory (mass, scalar-gravitational coupling constant, etc). In the previous section we found this coefficient for conformal scalar. In the following discussion it is more convenient to consider this coefficient to be some parameter of the theory. Then, the results are quite common and may be applied to arbitrary quantum bulk theory. We also suppose that there are no background bulk fields in the theory (except of bulk gravitational field).
Adding quantum bulk contribution to the action $`S`$ in (30) one can regard
$$S_{\mathrm{total}}=S+S_{\mathrm{Csmr}}$$
(53)
as the total action. In (52), $``$ is the radius of S<sub>4</sub> or H<sub>4</sub>.
In the bulk, one obtains the following equation of motion from $`S_{\mathrm{EH}}+S_{\mathrm{Csmr}}`$ by the variation over $`A`$:
$`0`$ $`=`$ $`\left(24_y^2A48(_yA)^2+{\displaystyle \frac{48}{l^2}}\right)\mathrm{e}^{4A}`$ (54)
$`+{\displaystyle \frac{1}{l^2}}\left(12_\sigma ^2A12(_\sigma A)^2+12\right)\mathrm{e}^{2A}+{\displaystyle \frac{16\pi Gc}{^5}}\mathrm{e}^A.`$
First, one can consider a special solution of the bulk equation (54). If one assumes that $`A`$ does not depend on $`\sigma `$, Eq.(54) has the following form:
$$0=\left(24_y^2A48(_yA)^2+\frac{48}{l^2}\right)\mathrm{e}^{4A}+\frac{16\pi Gc}{^5}\mathrm{e}^A.$$
(55)
Eq.(55) has the following integral:
$$E=\frac{1}{4}\left(\frac{d\left(\mathrm{e}^{2\stackrel{~}{A}}\right)}{dy}\right)^2+\frac{1}{l^2}\mathrm{e}^{4\stackrel{~}{A}}+\frac{1}{2l^2}\mathrm{e}^{2\stackrel{~}{A}}\frac{4\pi Gc}{3^5}\mathrm{e}^{\stackrel{~}{A}}.$$
(56)
Then if we assume $`\frac{d\left(\mathrm{e}^{2\stackrel{~}{A}}\right)}{dy}>0`$, we find the following solution in the bulk
$$y=\frac{1}{2}๐Q\left(E+\frac{Q^2}{l^2}+\frac{Q}{2l^2}\frac{4\pi Gc}{3^5}Q^{\frac{1}{2}}\right)^{\frac{1}{2}}.$$
(57)
This represents an example of self-consistent warped compactification. The analysis of such solution shows that for vanishing bulk cosmological constant it goes away from AdS space. This indicates that bulk Casimir effect acts against of warped compactification.
Let us discuss the solution in the situation when scale factor depends on both coordinates:$`y`$,$`\sigma `$. One can find the solution of (54) as an expansion with respect to $`\mathrm{e}^{\frac{y}{l}}`$ by assuming that $`\frac{y}{l}`$ is large:
$$\mathrm{e}^A=\frac{\mathrm{sinh}\frac{y}{l}}{\mathrm{cosh}\sigma }\frac{32\pi Gcl^3}{15^5}\mathrm{cosh}^4\sigma \mathrm{e}^{\frac{4y}{l}}+๐ช\left(\mathrm{e}^{\frac{5y}{l}}\right)$$
(58)
for the perturbation from the solution where the brane is S<sub>4</sub> and
$$\mathrm{e}^A=\frac{\mathrm{cosh}\frac{y}{l}}{\mathrm{sinh}\sigma }\frac{32\pi Gcl^3}{15^5}\mathrm{sinh}^4\sigma \mathrm{e}^{\frac{4y}{l}}+๐ช\left(\mathrm{e}^{\frac{5y}{l}}\right)$$
(59)
for the perturbation from H<sub>4</sub> brane solution.
On the brane at the boundary, one gets the following equation:
$`0`$ $`=`$ $`{\displaystyle \frac{48l^4}{16\pi G}}\left(_yA{\displaystyle \frac{1}{l}}\right)\mathrm{e}^{4A}+b^{}\left(4_\sigma ^4A16_\sigma ^2A\right)`$ (60)
$`4(b+b^{})\left(_\sigma ^4A+2_\sigma ^2A6(_\sigma A)^2_\sigma ^2A\right).`$
We should note that the contributions from $`S_{\mathrm{EH}}`$ and $`S_{\mathrm{GH}}`$ are twice from the naive values since we have two bulk regions which are connected with each other by the brane. Substituting the solutions (58) and (59) into (60), we find
$$0\frac{1}{\pi G}\left(\frac{1}{}\sqrt{1+\frac{^2}{l^2}}+\frac{64\pi Gl^7c}{3^{10}}\mathrm{cosh}^5\sigma \frac{1}{l}\right)^4+8b^{}.$$
(61)
for S<sub>4</sub> brane and
$$0\frac{1}{\pi G}\left(\frac{1}{}\sqrt{1+\frac{^2}{l^2}}+\frac{64\pi Gl^7c}{3^{10}}\mathrm{sinh}^5\sigma \frac{1}{l}\right)^4+8b^{}.$$
(62)
for H<sub>4</sub> brane. Here the radius $``$ of S<sub>4</sub> or H<sub>4</sub> is related with $`A(y_0)`$, if we assume the brane lies at $`y=y_0`$, by
$$\stackrel{~}{R}=l\mathrm{e}^{\stackrel{~}{A}(y_0)}.$$
(63)
In Eqs.(61) and (62), only the leading terms with respect to $`1/`$ are kept in the ones coming from $`S_{\mathrm{Csmr}}`$ (the terms including $`c`$). When $`c=0`$, the previous result in is reproduced. Eqs.(61) and (62) tell that the Casimir force deforms the shape of S<sub>4</sub> or H<sub>4</sub> since $``$ becomes $`\sigma `$ dependent. The effect becomes larger for large $`\sigma `$. In case of S<sub>4</sub> brane, the effect becomes large if the distance from the equator becomes large since $`\sigma `$ is related to the angle coordinate $`\chi `$ by (39). Especially at north and south poles ($`\chi =0`$, $`\pi `$), $`\mathrm{cosh}\sigma `$ diverges then $``$ should vanish as in Fig.1. Of course, the perturbation would be invalid when $`\mathrm{cosh}\sigma `$ is large. Thus, we demonstrated that bulk quantum effects do not destroy the quantum creation of de Sitter (inflationary) or Anti-de Sitter brane-world Universe. Of course, analytical continuation of 4d sphere to Lorentzian signature is supposed which leads to ever expanding inflationary brane-world Universe. However, as we see the bulk quantum effects change the effective radius of 4d sphere (or 4d hyperboloid).
When $`c=0`$, the solution can exist when $`b^{}<0`$ for S<sub>4</sub> brane (in this case it is qualitatively similar to quite well-known anomaly driven inflation of refs.) and $`b^{}>0`$ for H<sub>4</sub>. For S<sub>4</sub> brane, if $`b^{}<0`$, the effect of Casimir force makes the radius smaller (larger) if $`c>0`$ ($`c<0`$). For H<sub>4</sub> brane, from Eq.(62) for small $``$ it behaves as
$$0\frac{64l^7c}{3^{10}}\mathrm{sinh}^5\sigma +8b^{}.$$
(64)
Then one would have solution even if $`b^{}<0`$. (We should note $`\sigma `$ is restricted to be positive). Of course, one cannot do any quantitative conclusion since it is assumed $``$ is large when deriving (62).
We now compare the above obtained results with the case of no quantum correction $`W`$ (no matter) on the brane, i.e. when bulk quantum effects are leading. Putting $`b^{}=0`$ in (61) and (62), one gets
$$^8\frac{128\pi Gl^6c}{8}\mathrm{cosh}^5\sigma $$
(65)
for S<sub>4</sub> brane and
$$^8\frac{128\pi Gl^6c}{8}\mathrm{sinh}^5\sigma $$
(66)
for H<sub>4</sub> brane. Here we only consider the leading term with respect to $`c`$, which corresponds to large $``$ approximation. In case of S<sub>4</sub>, there is no real solution for negative $`b^{}`$ but there appears a solution for negative $`b^{}`$ in case of H<sub>4</sub>, where there is no solution without Casimir term $`S_{\mathrm{Csmr}}`$ in (52). Thus, we demonstrated that bulk quantum effects do not violate (in some cases ,even support) the quantum creation of de Sitter or Anti-de Sitter brane living in d5 AdS world.
## 4 Discussion
In summary, we compared the role of bulk matter and brane matter quantum effects in the realization of brane-world Universe with constant curvature 4d brane (de Sitter or Anti-de Sitter). In such Universe the bulk represents five-dimensional AdS space while observable four-dimensional Universe is the boundary (brane) of such five-dimensional space. The brane matter quantum effects may be included in the universal form, via the corresponding anomaly induced effective action on the brane. (Actually, they modify the effective brane tension which is fixed on classical level). In this way, the contribution of any specific conformally invariant matter theory (or even of brane quantum gravity) only changes the coefficients of the correspondent anomaly induced effective action. The bulk conformal matter quantum effects are more difficult to calculate. We made the correspondent evaluation for scalar in order to understand the qualitative structure of bulk effective action.
It is shown that quantum creation of inflationary (or hyperbolic) brane-world Universe is possible. Such Universe may be induced by only bulk quantum effects , or by only brane quantum effects. When both contributions are included the role of bulk quantum effects is in the deformation of shape of 5d AdS Universe as well as of shape of 4d spherical or hyperbolic brane. (These are induced by brane CFT quantum effects).
There are few possible extensions of the presented results. As we already mentioned in the introduction, it would be really interesting (and more consistent from the AdS/CFT correspondence point of view) to investigate the role of bulk quantum gravity in such self-consistent warped Randall-Sundrum compactification with curved boundary. This is under study currently.
From another side, one can combine the current scenario for quantum induced inflationary brane-world Universe with the ones where classical background matter is presented in the bulk and (or) on the brane. In any case, the number of possibilities to realize brane-world inflation occurs.
Acknowledgments. We would like to thank A. Bytsenko and K. Milton for helpful discussions. The work by SDO has been supported in part by CONACyT (CP, ref.990356 and grant 28454E) and in part by RFBR. The work by SZ and SDO has been supported in part by INFN, Gruppo Collegato di Trento.
|
warning/0006/math0006068.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Within the framework of the nonlinear DonnellโMushtariโVlasov (DMV) theory (see, e.g., ) the state of equilibrium of a transversely loaded thin isotropic elastic shell of uniform thickness is determined by the following system of two coupled nonlinear forth-order partial differential equations:
$$\begin{array}{c}D\mathrm{\Delta }^2w\epsilon ^{\alpha \mu }\epsilon ^{\beta \nu }w_{;\alpha \beta }\mathrm{\Phi }_{;\mu \nu }\epsilon ^{\alpha \mu }\epsilon ^{\beta \nu }b_{\alpha \beta }\mathrm{\Phi }_{;\mu \nu }=p,\hfill \\ (1/Eh)\mathrm{\Delta }^2\mathrm{\Phi }+(1/2)\epsilon ^{\alpha \mu }\epsilon ^{\beta \nu }w_{;\alpha \beta }w_{;\mu \nu }+\epsilon ^{\alpha \mu }\epsilon ^{\beta \nu }b_{\alpha \beta }w_{;\mu \nu }=0,\hfill \end{array}$$
(1)
in two independent variables, associated with the coordinates on the shell middle-surface $`F,`$ and two dependent variables โ the transversal displacement function $`w,`$ and Airyโs stress function $`\mathrm{\Phi }`$. Here, $`\epsilon ^{\alpha \beta }`$ is the alternating tensor of $`F`$; $`b_{\alpha \beta }`$ is the curvature tensor of $`F`$; $`D`$, $`E`$ and $`h`$ are the bending rigidity, Youngโs modulus and thickness of the shell, respectively (i.e., $`D`$, $`E`$ and $`h`$ are given constants); $`p`$ is the function of transversal load per unit surface area; a semicolon is used for covariant differentiation with respect to the metric tensor $`a_{\alpha \beta }`$ of the surface $`F`$; $`\mathrm{\Delta }`$ is the Laplace-Beltrami operator on $`F`$. Here and throughout, Greek indices have the range 1, 2 and the usual summation convention over a repeated index (one subscript and one superscript) is employed.
The present note is concerned with a special case of the nonlinear DMV theory; the so-called shallow shells are considered. In fact, this is an approximation of the theory, which is usually introduced as follows (cf., e.g. ). Let $`(x^1,x^2,x^3)`$ be a fixed right-handed rectangular Cartesian coordinate system in the 3-dimensional Euclidean space in which the middle-surface $`F`$ of a shell is embedded, and let this surface be given by the equation
$$x^3=f(x^1,x^2),(x^1,x^2)\mathrm{\Omega }๐^2,$$
where $`f:๐^2๐`$ is assumed to be a single-valued and smooth function possessing as many derivatives as may be required on the domain $`\mathrm{\Omega }`$. Let us take $`x^1,x^2`$ to serve as coordinates on the surface $`F`$. Then, relative to this coordinate system, the components of the fundamental tensors and the alternating tensor of $`F`$ are given by the expressions:
$$a_{\alpha \beta }=\delta _{\alpha \beta }+f_{,\alpha }f_{,\beta },b_{\alpha \beta }=a^{1/2}f_{,\alpha \beta },\epsilon ^{\alpha \beta }=a^{1/2}e^{\alpha \beta },$$
(2)
where
$$a=det(a_{\alpha \beta })=1+(f_{,1})^2+(f_{,2})^2;$$
$`\delta _{\alpha \beta }=\delta ^{\alpha \beta }`$ is the Kronecker delta symbol; $`e^{\alpha \beta }`$ is the alternating symbol; here and in what follows, a comma is used for partial differentiation with respect to the coordinates on $`F`$. A shell is said to be shallow on the domain $`\mathrm{\Omega }_0\mathrm{\Omega }`$ when the inequalities
$$\left|f_{,\alpha }\right|\left|f_{,\beta }\right|\epsilon ^21,\epsilon =const,$$
hold for every point $`(x^1,x^2)\mathrm{\Omega }_0`$. Hence, for shallow shells the quadratic terms in the right-hand sides of expressions (2) are small compared to unity and may be neglected. Thus, allowing for a relative error of order $`O\left(\epsilon ^2\right)`$, one may regard the intrinsic geometry of the shell middle-surface $`F`$ as Euclidean and $`(x^1,x^2)`$ may be thought of as an Euclidean coordinate system in which:
$`a_{\alpha \beta }`$ $`=`$ $`\delta _{\alpha \beta },`$ (3)
$`b_{\alpha \beta }`$ $`=`$ $`f_{,\alpha \beta },`$ (4)
$`\epsilon ^{\alpha \beta }`$ $`=`$ $`e^{\alpha \beta };`$ (5)
the mean curvature $`H`$ of the surface $`F`$ and its Gaussian curvature $`K`$ (note that the latter is not necessarily equal to zero within the allowed relative error) take the form
$`H`$ $`=`$ $`(1/2)\delta ^{\alpha \beta }f_{,\alpha \beta },`$ (6)
$`K`$ $`=`$ $`(1/2)e^{\alpha \mu }e^{\beta \nu }f_{,\alpha \beta }f_{,\mu \nu },`$ (7)
and system (1) reads
$$\begin{array}{c}D\delta ^{\alpha \beta }\delta ^{\mu \nu }w_{,\alpha \beta \mu \nu }e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }\mathrm{\Phi }_{,\mu \nu }e^{\alpha \mu }e^{\beta \nu }b_{\alpha \beta }\mathrm{\Phi }_{,\mu \nu }=p,\hfill \\ (1/Eh)\delta ^{\alpha \beta }\delta ^{\mu \nu }\mathrm{\Phi }_{,\alpha \beta \mu \nu }+(1/2)e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }w_{,\mu \nu }+e^{\alpha \mu }e^{\beta \nu }b_{\alpha \beta }w_{,\mu \nu }=0,\hfill \end{array}$$
(8)
Thus, we arrive at the equilibrium equations for shallow shells within the framework of the nonlinear DMV theory. Since equations (8) follow from Marguerreโs shell theory as well, they are also known as Marguerreโs equations for large deflection of plates with small initial curvature (i.e., shallow shells). These equations are well accepted and play an important role in the shell theory (see, e.g., , and the references therein). They also include as a special case, with $`b_{\alpha \beta }=0,`$ the well-known von Kรกrmรกn equations for large deflection of plates .
## 2 Symmetry Groups
The aim of the present work is to study, following , and , the invariance properties of system (8) relative to local one-parameter Lie groups of local point transformations acting on open subsets of the 4-dimensional Euclidean space $`๐^4`$, with coordinates $`(x^1,x^2,w,\mathrm{\Phi })`$, representing the involved independent and dependent variables. For that purpose Lie infinitesimal technique is used and, as a rule, the results obtained are expressed in terms of the infinitesimal generators (operators) of the groups; in the present case, the latter are vector fields of the form
$$๐=\xi ^\mu \frac{}{x^\mu }+\eta \frac{}{w}+\phi \frac{}{\mathrm{\Phi }},$$
(9)
where $`\xi ^\mu ,\eta `$ and $`\phi `$ are functions of the variables $`x^1,x^2,w`$ and $`\mathrm{\Phi }.`$ The system considered involves an arbitrary tensor field โ $`b_{\alpha \beta }`$, and an arbitrary function โ $`p`$. This gives rise to a group classification problem with respect to the arbitrary element โ the set $`\{b_{\alpha \beta },p\}`$.
The infinitesimal criterion of invariance leads to the following system of determining equations (DE system) for the components $`\xi ^1,\xi ^2,\eta `$ and $`\phi `$ of the vector fields of form (9) generating point symmetry groups admitted by system (8):
$$\xi ^1=C_1x^1+C_2x^2+C_3,$$
(10)
$$\xi ^2=C_2x^1+C_1x^2+C_4,$$
(11)
$$\eta =\eta (x^1,x^2),$$
(12)
$$\phi =B_1x^1+B_2x^2+B_3,$$
(13)
$$b_{\alpha \mu }\xi _{,\beta }^\mu +b_{\beta \mu }\xi _{,\alpha }^\mu +\xi ^\mu b_{\alpha \beta ,\mu }=\eta _{,\alpha \beta },$$
(14)
$$e^{\alpha \mu }e^{\beta \nu }b_{\alpha \beta }\eta _{,\mu \nu }=0,$$
(15)
$$D\delta ^{\alpha \beta }\delta ^{\mu \nu }\eta _{,\alpha \beta \mu \nu }=2p\xi _{,\mu }^\mu +\xi ^\mu p_{,\mu },$$
(16)
where $`C_1,\mathrm{},C_4,B_1,B_2,`$ and $`B_3`$ are arbitrary real constants. This result is established in through the standard computational procedure (see \[7, Sec. 5\] or \[8, Sec. 2.4\]), and therefore we can assert that the system (8) admits a vector field of form (9) if and only if (10) โ (16) hold. In other words, all possible symmetries of the kind under consideration inherent to system (8) can be found via the solution of the DE system. So, the problem to solve consists in finding the solutions of this system. The main difficulty here comes out of the determining equations (14) โ (16). They show that the space of solutions $`L`$ to the DE system, i.e., the Lie algebra of symmetries associated with system (8), depends on the choice of the arbitrary element $`\{b_{\alpha \beta },p\}`$. At this juncture, we do face a group classification problem. That is to say, that we will have to determine all those specializations (special forms) of the arbitrary element for which system (8) admits vector fields of the form (9).
Proceeding to analyze this problem, we will first show that equations (12), (14) โ (16) may be replaced by the following three equivalent ones:
$$\eta =\xi ^\mu f_{,\mu }+A_1x^1+A_2x^2+A_3,$$
(17)
where $`A_1,A_2`$ and $`A_3`$ are arbitrary real constants, and
$$2P\xi _{,\mu }^\mu +\xi ^\mu P_{,\mu }=0,$$
(18)
$$2K\xi _{,\mu }^\mu +\xi ^\mu K_{,\mu }=0,$$
(19)
where
$$P=2D\delta ^{\mu \nu }H_{,\mu \nu }+p.$$
(20)
Indeed, (17) represents the general solution of equations (14) of the form (12) when (4), (10) and (11) hold; under the same assumption, by substituting (17) into (15) and (16), and taking into account (6) and (7), after some algebra we obtain expressions (18) and (19). Using this result, hereafter we will assume that the DE system consists of equations (10), (11), (13), (17), (18) and (19).
Now, a 6-dimensional space of solutions to the DE system arises immediately. It corresponds to the 6-dimensional Lie algebra $`L_0`$ of the vector field whose components are given as follows
$`\xi ^\mu `$ $`=`$ $`0,`$
$`\eta `$ $`=`$ $`A_1x^1+A_2x^2+A_3,`$ (21)
$`\phi `$ $`=`$ $`B_1x^1+B_2x^2+B_3,`$
where $`A_1,A_2,A_3,B_1,B_2`$ and $`B_3`$ are arbitrary real constants. Evidently, the above solutions of the DE system do not depend on the choice (specialization) of the arbitrary element. Hence, for any specialization of the arbitrary element system (8) admits the 6-parameter Lie group $`G_0`$ generated by all linear combinations of the vector fields
$$\frac{}{w},x^1\frac{}{w},x^2\frac{}{w},\frac{}{\mathrm{\Phi }},x^1\frac{}{\mathrm{\Phi }},x^2\frac{}{\mathrm{\Phi }};$$
these vector fields constitute a basis of the associated Lie algebra $`L_0`$. Simultaneously, $`G_0`$ is the largest Lie group of point transformations admitted by system (8) for any choice of the arbitrary element. Indeed, by setting
$$f=\mathrm{sin}(x^1)\mathrm{sin}(x^2),$$
for instance, one can easily verify that for this particular choice of the arbitrary element the DE system has no other solutions besides those given by (21).
So far, we have obtained the so-called kernel of the full symmetry groups associated with system (8), that is the group $`G_0`$. The next step is to identify the cases in which the system under consideration possesses larger groups of point symmetries. In the light of all the above, it means to characterize in a suitable manner all those specializations of the arbitrary element for which system (8) admits vector fields of the form
$$๐=\xi ^\mu \frac{}{x^\mu }\xi ^\mu f_{,\mu }\frac{}{w}(\xi ^\mu 0),$$
(22)
where $`\xi ^\mu `$ are given by the expressions (10) and (11), keeping in mind that (18) and (19) remain the only conditions (necessary and sufficient) for (8) to be invariant under a group generated by a vector field of form (22). Note that all determining equations are thus taken into account.
Taking into account (6), (7) and (20) we can see at once that for $`f=p=0`$ the invariance conditions (18) and (19) are satisfied. Hence, system (8), with $`f=p=0`$, possesses a larger group of point symmetries (in addition to $`G_0`$) and (10), (11) and (22) show that this is the complete 4-parameter group of homothetic motions of the Euclidean plane. In this special case, (8) coincide with the homogeneous von Kรกrmรกn equations, so that we have arrived at the result obtained in (see also ). This example gives us a good motivation for studying the general case.
We begin with the following observation. Given a vector field $`๐`$of form (22), the function
$$\stackrel{~}{w}=w+f,$$
(23)
is an invariant of the corresponding Lie group of transformations acting on $`๐^4`$. Therefore, we can introduce new coordinates $`(x^1,x^2,\stackrel{~}{w},\mathrm{\Phi })`$ on $`๐^4`$, the new dependent variable $`\stackrel{~}{w}`$ being defined by (23), in which $`๐`$ takes the form of an infinitesimal operator of the Lie algebra associated with the group of homothetic motions of the Euclidean plane, namely
$$๐=\xi ^\mu \frac{}{x^\mu },$$
(24)
(note that the components $`\xi ^1`$ and $`\xi ^2`$ still have the form (10) and (11), respectively, as before), and system (8) reads
$$\begin{array}{c}D\delta ^{\alpha \beta }\delta ^{\mu \nu }\stackrel{~}{w}_{,\alpha \beta \mu \nu }e^{\alpha \mu }e^{\beta \nu }\stackrel{~}{w}_{,\alpha \beta }\mathrm{\Phi }_{,\mu \nu }=P,\hfill \\ (1/Eh)\delta ^{\alpha \beta }\delta ^{\mu \nu }\mathrm{\Phi }_{,\alpha \beta \mu \nu }+(1/2)e^{\alpha \mu }e^{\beta \nu }\stackrel{~}{w}_{,\alpha \beta }\stackrel{~}{w}_{,\mu \nu }=K.\hfill \end{array}$$
(25)
Thus, the problem of invariance of system (8) with respect to a vector field (22) converts into the problem of invariance of system (25) under a vector field (24) as a change of the variables does not affect the group properties of a system of differential equations. Now, taking into account the invariance conditions (18) and (19) (note that they remain unchanged under the above coordinate transformation), we can conclude that system (25) admits a one-parameter group $`G`$ of homothetic motions of the Euclidean plane generated by a vector field of form (24) if and only if the corresponding arbitrary element $`\{b_{\alpha \beta },p\}`$ is such that the functions $`P`$ and $`K`$, defined by formulae (7) and (20), are invariants of $`G`$ (when $`C_1=0`$) or eigenfunctions (when $`C_10`$) of its generator, the latter being regarded as an operator acting on the smooth functions $`\zeta :MR,M๐^2`$. This result may be thought of as a general solution to the group classification problem under consideration in terms of the function $`p`$ and the two characteristic invariants, $`H`$ and $`K`$, of the shell middle-surface $`F`$.
## 3 Equivalence Transformations
Let us now briefly discuss, in the context of the shell theory, the meaning of the coordinate transformation $`\omega :๐^4๐^4,(x^1,x^2,w,\mathrm{\Phi })(x^1,x^2,w+f,\mathrm{\Phi })`$ introduced in the previous Section. Omitting tildeโs in equations (25), we can say that systems (8) and (25) belong to the same class โ they have the same differential structure, and differ from one another only in the form of the arbitrary element. Hence, according to \[7, Definition 6.4\], we may conclude that $`\omega :๐^4๐^4`$ is an equivalence transformation for system(8). In addition, we see that (25) is nothing but a system of nonhomogeneous von Kรกrmรกn equations with special right-hand sides. It is noteworthy that Marguerreโs equations (8), i.e., the equilibrium equations for shallow shells, turned out to be equivalent to the von Kรกrmรกn equations, i.e., the equilibrium equations for plates, regardless of the invariance properties of system (8). To the best of the authorโs knowledge, this fact has not been noticed before in the literature, though both the von Kรกrmรกn equations and Marguerreโs equations have been studied and utilized for many years until now (see, e.g., , , , , , , and the references therein). In our opinion, the established correspondence between these two systems of equations will certainly be of use in solving a wide range of problems arising in the shell theory. In particular, the results for the von Kรกrmรกn equations obtained in can be easily conveyed to the theory of shallow shells.
Several additional applications of the transformation $`\omega `$ can by given too. Evidently, it maps:
$``$ the time-dependent Marguerreโs equations
$$\begin{array}{c}D\delta ^{\alpha \beta }\delta ^{\mu \nu }w_{,\alpha \beta \mu \nu }e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }\mathrm{\Phi }_{,\mu \nu }e^{\alpha \mu }e^{\beta \nu }b_{\alpha \beta }\mathrm{\Phi }_{,\mu \nu }+\rho \ddot{w}=p,\hfill \\ (1/Eh)\delta ^{\alpha \beta }\delta ^{\mu \nu }\mathrm{\Phi }_{,\alpha \beta \mu \nu }+(1/2)e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }w_{,\mu \nu }+e^{\alpha \mu }e^{\beta \nu }b_{\alpha \beta }w_{,\mu \nu }=0,\hfill \end{array}$$
(where $`\rho `$ is the mass per unit area of the shell middle-surface, and a superposed dot is used to denote partial derivative with respect to the time $`t`$) into the time-dependent von Kรกrmรกn equations
$$\begin{array}{c}D\delta ^{\alpha \beta }\delta ^{\mu \nu }w_{,\alpha \beta \mu \nu }e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }\mathrm{\Phi }_{,\mu \nu }+\rho \ddot{w}=P,\hfill \\ (1/Eh)\delta ^{\alpha \beta }\delta ^{\mu \nu }\mathrm{\Phi }_{,\alpha \beta \mu \nu }+(1/2)e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }w_{,\mu \nu }=K;\hfill \end{array}$$
$``$ the time-dependent equations for anisotropic shallow shells
$$\begin{array}{c}D^{\alpha \beta \mu \nu }w_{,\alpha \beta \mu \nu }e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }\mathrm{\Phi }_{,\mu \nu }e^{\alpha \mu }e^{\beta \nu }b_{\alpha \beta }\mathrm{\Phi }_{,\mu \nu }+\rho \ddot{w}=p,\hfill \\ E^{\alpha \beta \mu \nu }\mathrm{\Phi }_{,\alpha \beta \mu \nu }+(1/2)e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }w_{,\mu \nu }+e^{\alpha \mu }e^{\beta \nu }b_{\alpha \beta }w_{,\mu \nu }=0,\hfill \end{array}$$
(where $`D^{\alpha \beta \mu \nu }`$ and $`E^{\alpha \beta \mu \nu }`$ denote the material constants) into the following system of von Kรกrmรกn-type equations
$$\begin{array}{c}D^{\alpha \beta \mu \nu }w_{,\alpha \beta \mu \nu }e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }\mathrm{\Phi }_{,\mu \nu }+\rho \ddot{w}=D^{\alpha \beta \mu \nu }f_{,\alpha \beta \mu \nu }+p,\hfill \\ E^{\alpha \beta \mu \nu }\mathrm{\Phi }_{,\alpha \beta \mu \nu }+(1/2)e^{\alpha \mu }e^{\beta \nu }w_{,\alpha \beta }w_{,\mu \nu }=K.\hfill \end{array}$$
The above list could be continued.
Acknowledgments
This research was supported by Contract No. MM 517/1995 with the National Scientific Fund, Republic of Bulgaria.
|
warning/0006/hep-ph0006097.html
|
ar5iv
|
text
|
# 1 *
The existence of extra spacetime dimensions has recently been suggested as a means to explain the hierarchy. In one scenario of this kind from Arkani-Hamed, Dimopoulos, and Dvali (ADD), the apparent hierarchy is generated by a large volume for the extra dimensions. In this case, the fundamental Planck scale in $`4+n`$-dimensions, $`M`$, can be reduced to the TeV scale and is related to the observed 4-d Planck scale, $`\overline{M}_{Pl}`$, through the volume $`V_n`$ of the compactified dimensions, $`\overline{M}_{Pl}^2=V_nM^{2+n}`$. In a second scenario due to Randall and Sundrum (RS), the observed hierarchy is induced through an exponential warp factor which arises from a non-factorizable geometry. An exciting feature of these approaches is that they both lead to concrete and distinctive phenomenological tests at the TeV scale.
In addition to collider tests, loop-order processes, such as rare transitions which are suppressed in the Standard Model (SM) or radiative corrections to perturbatively calculable processes, can provide complementary information about new physics. One such traditional quantity is the $`(g2)`$ of the $`\mu `$. Currently the SM prediction is approximately $`1\sigma `$ higher than that of the World Average measured value, with the difference between the theoretical and experimental results being $`a_\mu ^{exp}a_\mu ^{SM}=(43\pm 45)\times 10^{10}`$, where $`a=(g2)/2`$. This corresponds to a $`95\%`$ CL upper bound on the magnitude of a new negative contribution, $`\mathrm{\Delta }a_\mu `$, of $`3.1\times 10^9`$. The E821 experiment at BNL is expected to reduce the experimental error on $`a_\mu ^{exp}`$ by approximately an order of magnitude during the next few years to the level of 0.35ppm which is below the current SM theory error of 0.60ppm. The SM error will also decrease in the future as more data on the $`R`$ ratio in the low energy region becomes available. The size of the contribution to $`(g2)_\mu `$ in the ADD scenario has been calculated in Ref. and results in interesting constraints. In this paper, we examine this quantity within the RS model in the case where the SM fields propagate in the bulk with the expectation from our earlier work that existing data will yield interesting bounds over a region of the parameter space. With an anticipated ten-fold increase in the experimental precision in the not too distant future, these bounds should soon improve if no signal for new physics is observed.
In its original construction, the RS model consists of two 3-branes each being stabilized at an $`S^1/Z_2`$ orbifold fixed point with a separation of $`\pi r_c`$ between the branes in an additional dimension denoted as $`r_c\varphi `$. The model initially postulated that only gravity was allowed to propagate in the higher dimensional anti-deSitter bulk with the SM fields being confined to one of the 3-branes. The exponential โwarpโ factor $`e^{kr_c\varphi }`$, with $`k`$ being a 5-d space-time curvature parameter of order the Planck scale, is responsible for generating the observed hierarchy assuming the scale of physics on the SM brane located at $`\varphi =\pi `$ is $`\mathrm{\Lambda }_\pi =\overline{M}_{Pl}e^{kr_c\pi }1`$ TeV with $`kr_c1112`$. The usual 4-d Planck scale and that of the original 5-d theory are found to be related via $`\overline{M}_{Pl}^2=M_5^3/k`$. Recently, a series of authors have considered peeling the SM gauge and matter fields off of the wall in the limit where their back-reaction on the RS metric can be ignored. (There are a number of arguments which strongly suggest that if the Higgs is the source of electroweak symmetry breaking it must remain on the wall.) It is the existence of these SM bulk fields that allows for a potentially sizeable contribution to $`(g2)_\mu `$. In what follows we use the notation as defined in the last paper listed in Ref..
When the SM gauge and matter fields are allowed to propagate in the bulk, there are three parameters that need to be specified to determine the phenomenological predictions of the RS model: $`ck/\overline{M}_{Pl}`$ which is expected to lie in the range 0.01 to 1, the common dimensionless bulk mass parameter for the fermions $`\nu m/k`$, where $`m`$ represents the 5-d fermion mass, which is expected to be of order unity, and the mass of the lightest gauge, fermion or graviton Kaluza-Klein(KK) excitation. We remind the reader that a common value of $`\nu `$ for all fermions is not a necessary assumption but is certainly the simplest choice and the one which naturally avoids constraints associated with flavor changing neutral currents. For a fixed value of $`\nu `$,the entire KK spectrum is determined for all fields once the mass of a single KK excitation is known. We recall that the KK spectrums for gravitons, fermions, and gauge bosons are related by the roots of various Bessel functions and that all gauge bosons, i.e., gluons, $`W`$โs, $`Z`$โs and $`\gamma `$โs, have essentially the same excitation spectra.
Note that the limits we obtain below are derived under the assumption that no other new physics is present beyond what is considered here. As with all bounds obtained via indirect means, the presence of additional new interactions may cancel the loop effects and erase or ease the constraints.
Consider the situation where we have two fermions in the bulk, $`D`$ and $`S`$, which have the quantum numbers of an $`SU(2)_L`$ doublet and singlet with weak hypercharges $`Y=1/2`$ and $`1`$, respectively. Following the notation of our previous work, their interactions with the gauge fields can be described by the action,
$$S_{fV}=d^4xr_cd\varphi \sqrt{G}[V_n^M(\frac{i}{2}\overline{S}\gamma ^n๐_MS+h.c.)sgn(\varphi )m_S\overline{S}S+(SD)],$$
(1)
where, $`G`$ is the determinant of the metric tensor, $`V`$ is the vielbein, $`๐_M`$ is a covariant derivative and $`h.c.`$ denotes the Hermitian conjugate term. Here, as discussed above, we will assume that $`m_D=m_S=k\nu `$. Note that gauge interactions do not mix the $`D`$ and $`S`$ fields. The $`D`$ and $`S`$ fields also interact with the Higgs isodoublet field(s), $`H^0`$, which reside on the wall, i.e.,
$$S_{fH}=\frac{\lambda }{k}d^4x๐\varphi \sqrt{G}\overline{S}DH^0\delta (\varphi \pi )+h.c.,$$
(2)
with $`\lambda `$ being a dimensionless Yukawa coupling. Due to the KK mechanism the fields $`D_{L,R}^{(n)}`$ and $`S_{L,R}^{(n)}`$ form separate 4-d towers of Dirac fermions which are degenerate level by level. The KK expansion can be written as $`D=D_L^{(n)}(x)\chi ^{(n)}(\varphi )+D_R^{(n)}(x)\tau ^{(n)}(\varphi )`$ and $`S=S_L^{(n)}(x)\tau ^{(n)}(\varphi )+S_R^{(n)}(x)\chi ^{(n)}(\varphi )`$ where the $`\nu `$-dependent $`\chi (\tau )`$ fields are $`Z_2`$ even(odd) and are given explicitly in our previous paper. Note that the $`Z_2`$ orbifold symmetry allows couplings of the type $`\overline{D_L}^{(n)}S_R^{(m)}+h.c.`$ but not ones of the form $`\overline{D_R}^{(n)}S_L^{(m)}+h.c.`$ since the $`Z_2`$ odd wavefunctions vanish on both of the boundaries. After shifting the Higgs field $`H^0e^{kr_c\pi }H`$ so that it is canonically normalized, the value of $`\lambda `$ is fixed as a function of $`\nu `$ by the requirement that the coupling of the $`D_L`$ and $`S_R`$ zero modes obtains a mass, $`m_\mu `$, once the Higgs gets a vev, $`v_4=v_{SM}/\sqrt{2}`$ with $`v_{SM}246`$ GeV. (It is thus important to observe that in a theory with a fixed value of $`\nu `$ the set of Yukawa couplings $`\lambda _f`$ associated with the set of SM fermions is clearly hierarchical.) This then fixes the couplings between a Higgs, a zero mode fermion and any tower member to be $`C_{0i}^{ffH}=(1)^i\frac{m_\mu }{v_4}\sqrt{F}`$, as well as the coupling between two different tower members and the Higgs as $`C_{ij}^{ffH}=(1)^{i+j}\frac{m_\mu }{v_4}F`$ (up to a possible $`\nu `$-dependent sign where),
$$F=2\left|\frac{1ฯต^{2\nu +1}}{1+2\nu }\right|,$$
(3)
with $`ฯต=e^{kr_c\pi }10^{16}`$. Note that for negative values of $`\nu `$, the factor $`F`$ grows exponentially large.
In terms of the $`D`$ and $`S`$ fields, the operator which generates the anomalous magnetic dipole moment of the $`\mu `$ can be written as $`D_L^{(0)}\sigma _{\mu \nu }S_R^{(0)}+h.c.`$. This reminds us that this operator and the muon mass generating term have the same isospin and helicity structure such that a Higgs interaction is required in the form of a mass insertion to connect the two otherwise decoupled zero modes. We can think of this mass insertion as the interaction of a fermion with an external Higgs field that has been replaced by its vev.
Helicity flips play an important role in evaluating the contributions to $`(g2)_\mu `$ since muon KK excitations are now propagating inside the loop. As is well-known, for non-chiral couplings the contribution to the anomalous magnetic moment of a light fermion can be enhanced when a heavy fermion of mass $`m_h`$ participates inside the loop. There are a number of diagrams that can contribute to $`(g2)_\mu `$ at one loop of which two are shown in Fig. 1. The diagram on the left corresponds to the exchange of a tower of the 4-d neutral gauge bosons, $`\gamma ^{(n)}`$ and/or $`Z^{(n)}`$, which we will now discuss in detail. Due to gauge invariance we are free to choose a particular gauge in order to simplify the calculation. Here, we make use of the $`\xi =1`$ unitary gauge where the 4-d propagator is just the flat space metric tensor. Hence, the loops with the 4-d components of the gauge fields and the ones with the fifth component need to be considered separately. In this example, the mass insertion takes place inside the loop before the photon is emitted. Clearly there are three other diagrams of this class: two with the mass insertion on an external leg and the third with the mass insertion inside the loop but after the photon is emitted. The amplitude arising from this vector exchange graph is given by
$`_V`$ $`=`$ $`C_LC_RC_{0in}^{ffA}C_{0jn}^{ffA}\overline{u}(p^{})(ie\gamma _\mu )P_Li{\displaystyle \frac{\widehat{\overline{)}p^{}}+m_j}{\widehat{p}^2m_j^2}}(ie\gamma _\nu )P_L`$ (4)
$`\times `$ $`i{\displaystyle \frac{\widehat{\overline{)}p}+m_j}{\widehat{p}^2m_j^2}}(im_{ij}P_R)i{\displaystyle \frac{\widehat{\overline{)}p}+m_i}{\widehat{p}^2m_i^2}}(ie\gamma ^\mu )P_Ru(p){\displaystyle \frac{i}{k^2m_A^{(n)2}}}+h.c.,`$
where $`C_{L,R}`$ are the corresponding couplings of the SM gauge boson to the $`\mu `$ in units of $`e`$ and $`\widehat{p}(\widehat{p}^{})=p(p^{})k`$. The coefficients $`C_{0in}^{ffA}`$ are the reduced couplings between a zero mode fermion, a fermion tower member of mass $`m_i`$ and the $`n^{th}`$ gauge boson tower member. Here $`m_i`$ and $`m_j`$ are the masses of the $`D`$ or $`S`$ fermionic KK states and $`m_A^{(n)}`$ are the masses of the KK gauge tower states. Note that the mass insertion, $`m_{ij}`$, comes with a chirality factor that can be determined from the action $`S_{fH}`$; numerically, $`m_{ij}=C_{ij}^{ffH}v_4`$. The amplitude where the mass insertion comes after the photon emission can be easily obtained by interchanging $`i`$ and $`j`$ in the resulting final amplitude expression. When the mass insertion occurs on an external leg it connects a zero mode with a tower mode and is given by $`m_{0i}=C_{0i}^{ffH}v_4`$. With some algebra it is straightforward to show that the corresponding amplitudes obtained in the two cases with external insertions are suppressed in comparison to the case of internal insertion by a factor of order $`m_\mu ^2/M_{KK}^2`$, where $`M_{KK}`$ is a typical large KK mass. In the case of the $`W`$ gauge boson tower graphs, since the $`W`$ couples only to the $`D`$โs, the mass insertion must occur on the incoming leg of the graph and the photon is emitted from the $`W`$; this graph can also be shown to produce a sub-leading contribution by a factor of order $`m_\mu ^2/M_{KK}^2`$. Thus, $`W`$ tower graphs can be safely ignored in comparison to those arising from the $`Z`$ and $`\gamma `$ towers and the resulting contribution from all of the 4-d vector exchanges, neglecting the subleading terms, is given by
$`\mathrm{\Delta }a_\mu ^A`$ $`=`$ $`{\displaystyle \underset{i,j,k}{}}{\displaystyle \frac{3C_V^2K\alpha m_{ij}m_\mu }{\pi }}\times `$ (5)
$`{\displaystyle _0^1}๐x{\displaystyle _0^{1x}}๐y\left[{\displaystyle \frac{x(1xy)}{m_j^2(1xy)+m_i^2y+m_A^{(k)2}x}}+(ij)\right],`$
where the sum is over all internal KK states, the $`ij`$ represents the addition of the other internal insertion graph and $`C_V^2=C_{0ik}^{ffA}C_{0jk}^{ffA}`$ with $`K`$ given by
$$K=\left[1+\frac{(14x_w)^21}{16x_w(1x_w)}\right],$$
(6)
where $`x_w=\mathrm{sin}^2\theta _w0.2315`$. Note that since the coefficients $`C_{0ik}^{ffA}`$ behave as $`1/\sqrt{F}`$ and $`m_{ij}F`$ we may expect that the $`\nu `$ dependence of $`\mathrm{\Delta }a_\mu ^A`$ to be rather weak. In principle the sum extends over all of the internal KK states but in practice we find that truncating the sum after the first 20-40 members of each tower leads to a rather stable result.
The next class of graphs is similar to the 4-d vector exchange, but in the $`\xi =1`$ gauge, now involves the fifth component of the original 5-d field. Here it is important to recall that these fifth components are $`Z_2`$ odd fields thus connecting $`S_L(D_L)`$ with $`S_R(D_R)`$. The action $`S_{fH}`$ in Eq.(2) demonstrates that the Higgs boson does not interact with odd fields since they vanish on the wall. From this observation we conclude that in the case of fifth component vector exchanges the mass insertions can only occur on the external legs. By following similar algebraic manipulations as before it is easy to show that all of these contributions are always subleading by factors of order $`m_\mu ^2/M_{KK}^2`$ and thus their contribution to $`(g2)_\mu `$ can be safely neglected.
Next, we turn to the possibility of Higgs exchange, also shown in Fig. 1. Ordinarily, one might dismiss such contributions as being small but they now involve the off-diagonal Higgs couplings discussed above which contain powers of the factor $`F`$ which grows large rapidly as $`\nu `$ grows negative. The amplitude for the Higgs graph shown in the figure is given by
$`_H`$ $`=`$ $`C_{0i}^{ffH}C_{0j}^{ffH}\overline{u}(p^{})iP_Ri{\displaystyle \frac{\widehat{\overline{)}p^{}}+m_i}{\widehat{p}^2m_i^2}}(ie\gamma _\nu )P_Ri{\displaystyle \frac{\widehat{\overline{)}p}+m_i}{\widehat{p}^2m_i^2}}`$ (7)
$`\times `$ $`(im_{ij}P_L)i{\displaystyle \frac{\widehat{\overline{)}p}+m_j}{\widehat{p}^2m_j^2}}iP_Ru(p){\displaystyle \frac{i}{k^2m_H^2}}+h.c.,`$
using the notation above. Here the coefficients $`C_{0i}^{FFH}`$ are the reduced couplings of a Higgs boson to a zero mode fermion and an $`i^{th}`$ fermion tower member. The amplitude where the insertion and emission occur with the opposite order can be obtained in a straightforward manner and, as we now expect, the two diagrams with external insertions can be shown to be subleading. Combining the two dominant Higgs amplitudes we find the following contribution to $`\mathrm{\Delta }a_\mu `$:
$$\mathrm{\Delta }a_\mu ^H=\underset{i,j}{}\frac{C_H^2m_{ij}m_\mu }{16\pi ^2}_0^1๐x_0^{1x}๐y\left[\frac{(1xy)(3y3x1)}{m_i^2(1xy)+m_j^2y+m_H^2x}+(ij)\right],$$
(8)
where $`C_H^2=C_{0i}^{ffH}C_{0j}^{ffH}`$ with $`m_H`$ being the Higgs boson mass. The $`ij`$ term results from the addition of the other internal emission graph. (In our numerical analysis below we assume $`m_H=120`$ GeV; these results are not very sensitive to this particular choice.) Since $`C_{0i}^{ffH}`$ behave as $`\sqrt{F}`$ and $`m_{ij}F`$ we expect $`\mathrm{\Delta }a_\mu ^H`$ to have a strong $`\nu `$ dependence and to grow very rapidly as $`\nu `$ becomes increasingly negative. As in the vector case, truncating the sum over the KK fermion contributions after the first 20-40 tower members have been included yields a numerically stable result.
What are the other potential contributions to $`(g2)_\mu `$ in the RS model? The radion is the zero mode remnant scalar resulting from the KK decomposition of the 5-d graviton field. Since it couples diagonally to KK tower members, as does the zero mode graviton and photon, the $`\mu `$ continues through the diagram and no KK modes are excited. Loops involving radions are thus easily shown to be small since both the radion couplings and the mass insertions in this case are not accompanied by any compensating powers of $`F`$. These contributions can be safely neglected.
The last remaining potential contribution arises from graviton loops which may be calculated via the Feynman rules given in with small modifications due to the fact that the SM fields are now in the bulk and have nontrival $`Z_2`$ parity. These diagrams lead to amplitudes which are found to be log-divergent and lead to cutoff ($`\mathrm{\Lambda }`$) dependent results and are thus not well-defined. The divergences, which occur in all graviton diagrams, arise due to the fact that the operator describing, e.g., the fermion-fermion-graviton interaction is dimension-five and involves an additional power of fermion momentum.
This differs significantly from the results obtained by Graesser in the case of the ADD model, where it was found that the total contribution to $`(g2)_\mu `$ due to gravity is finite. This difference arises from a number of sources: (i) in the ADD model the SM fields lie on the wall and only gravitons are allowed to propagate in the bulk, whereas in the version of the RS model under consideration here, both the SM gauge fields and fermions propagate in the bulk. (ii) In the ADD case, the graviton couplings are universal for all KK tower members whereas in the RS scenario the couplings are KK excitation state dependent and also differ for fermions and gauge bosons for arbitrary values of $`\nu `$. (iii) In the ADD case, each of the 5 diagrams shown in Fig. 1 of Graesser was found to be log divergent with their sum, however, being finite, since the divergences cancel at each KK level. In the RS case, these cancellations cannot occur due to both the breakdown in universality of the graviton couplings and the fact that the complete calculation of each diagram involves different coupling coefficients and different numbers of KK states. For example, in the diagram where the graviton is emitted off the fermion KK line, we must sum over the triple product of coefficients $`C_{0ik}^{ffG}C_{0jk}^{ffG}m_{ij}`$. On the otherhand, the diagram involving the gauge-gauge graviton vertex we must instead sum over the quartic product of coefficients $`C_{0ik}^{ffA}C_{0k\mathrm{}}^{AAG}C_{0n\mathrm{}}^{ffG}m_{in}`$, where the sum extends over the KK towers for two fermions, one graviton, and one photon. Since the evaluation of the coefficients involve $`\nu `$-dependent integrals over Bessel functions, it is highly unlikely that the divergences encountered in each class of diagrams can sum to zero unless a theorem demands that it be so. Thus, we expect that the complete contribution due to gravity to remain log divergent when all diagrams are summed.
We can, however, make an estimate of the size of these graviton contributions. As in the case of vector and Higgs boson exchange we expect graphs with internal insertions proportional to $`F`$ to dominate. Since the relevant vertex couplings $`C_{0ik}^{ffG}`$ scale as $`1/\sqrt{F}`$ we do not expect the graviton contribution to be strongly $`\nu `$ dependent as was also the case for vector exchange. An order of magnitude estimate suggests that
$$\mathrm{\Delta }a_\mu ^G\underset{ijk}{}\frac{C_{0ik}^{ffG}C_{0jk}^{ffG}m_{ij}m_\mu }{16\pi ^2\mathrm{\Lambda }_\pi ^2}log\left[\frac{\mathrm{\Lambda }^2}{m_{KK_k}^2}\right],$$
(9)
with the cutoff $`\mathrm{\Lambda }`$ expected to be of order $`\mathrm{\Lambda }_\pi `$. The origin of the various terms in this estimate are easy to identify: the $`C`$โs are vertex functions, the $`(16\pi ^2)^1`$ is a typical loop factor, the $`\mathrm{\Lambda }_\pi ^1`$ appears in all graviton couplings to SM fields, and the $`m_{ij}m_\mu `$ arises in the usual single mass insertion approximation as seen above. The $`\mathrm{log}\mathrm{\Lambda }^2`$ is the divergence discussed above and $`m_{KK_i}`$ represents canonical KK-tower masses which appear in the loop. Note that $`\mathrm{\Lambda }_\pi \mathrm{}`$ as $`m_{KK_i}\mathrm{}`$ and thus $`\mathrm{\Delta }a_\mu ^G0`$ as the KK mass gets large. Inserting typical values of the parameters we estimate that $`\mathrm{\Delta }a_\mu ^G`$ should certainly be less than $`10^{(910)}`$ and thus is at most comparable to the vector boson contribution. (An explicit calculation of the diagram containing the $`\gamma ffG`$ vertex confirms this expectation.) As we will see this means that the graviton exchange contributions will then have very little effect upon our results.
Let us now turn to our numerical results which are summarized in Fig. 2. We first remind the reader that for a fixed value of $`\nu `$ specifying the mass of any single KK tower state determines the entire KK spectrum for fermions, gauge bosons and gravitons. Here we take the mass of the lightest gauge KK state to be 1 TeV with the results for both the vector and Higgs contributions scaling as $`(1\mathrm{TeV}/M_{KK})^2`$. In our previous work we have shown that for values of $`\nu \begin{array}{c}>\hfill \\ \hfill \end{array}0.3`$ the masses of the KK states as well as $`\mathrm{\Lambda }_\pi `$ are required to be in the multi-TeV range, disfavoring this model as a solution to the hierarchy problem. Also we noted that when $`\nu \begin{array}{c}<\hfill \\ \hfill \end{array}0.9`$ the Yukawa couplings of the fermions became too large, hence for we display the range of $`\nu `$ in Fig. 2 to be between $`0.3`$ and $`1`$.
The first result to notice is the value of $`\mathrm{\Delta }a_\mu ^A`$ corresponding to the essentially $`\nu `$-independent horizontal dotted curve. This $`\nu `$ independence is only approximate and results from the scale of the figure as $`\mathrm{\Delta }a_\mu ^A`$ varies by a factor of order a few as $`\nu `$ increases up to 2. This curve lies about an order of magnitude below the current experimental $`(g2)_\mu `$ bound implying that we should obtain no reasonable constraint on the KK masses arising from this contribution unless there is a substantial change in both the experimental central value and the error. The rising dotted line is the Higgs boson contribution which we note is rather small over most of the interesting range of $`\nu `$. Scaling the curve to allow for the first KK masses to only be as large as a few to 10 TeV, we find that the region to the left of $`\nu 0.70`$ is excluded. This rules out a reasonable fraction, $`30\%`$, of the preferred allowed range of $`\nu `$ remaining subsequent to our last analysis.
There are however two other points to consider. First, a quick examination of the perturbative bound on the Yukawa couplings in the Higgs graph shown in Fig. 1 also yields a constraint. Imposing the weak requirement that $`C_H^2=C_{0i}^{ffH}C_{0j}^{ffH}<16\pi ^2`$ we obtain the dash-dotted line in Fig. 2 at $`\nu 0.77`$ excluding the region to its left. A second, perturbative-like bound can also be extracted from the Higgs loop in Fig.1 when we remove the photon line and consider the resulting mass renormalization contribution. We next demand the no fine-tuning requirement that the finite part of this graph be not much larger than $`m_\mu `$; from this we explicitly obtain the constraint that $`C_H^2m_{ij}/16\pi ^2`$ be not much greater than $`m_\mu `$. This then excludes the region to the left of the solid line at $`\nu 0.64`$ and results in a stronger bound than that obtained from the present experimental value of $`(g2)_\mu `$. Of course one could repeat this exercise for the case of the top quark provided the value of $`\nu `$ is universal. In this case, the value of $`C_H^2m_{ij}`$ is increased by the factor $`(m_t/m_\mu )^2`$ and the resulting bound is drastically strengthened. We find that the region to the left of $`\nu 0.44`$ would now be excluded by this analysis; these considerations now exclude more than $`75\%`$ of the previously preferred range of $`\nu `$ and leaves only the relatively narrow window between $`0.30`$ and $`0.44`$ as allowed. This would seem to greatly disfavor the possibility of the SM being in the RS bulk in the case of a universal mass parameter, $`\nu `$.
In conclusion the present experimental measurements of $`(g2)_\mu `$ do not place significant constraints on localized gravity models with the SM field content in the bulk for most of the allowed range of $`\nu `$. However, requiring that one loop corrections to the fermion masses be of the same order as the measured fermion mass, so that no fine tuning of the parameters in the Lagrangian are necessary, severly constrains the allowed bulk mass parameter space. This result is obtained assuming that the value of $`\nu `$ is flavor independent and that the 5-d Yukawa couplings of the fermions are hierarchical. Given these assumptions, however, placing the SM field content in the bulk is tightly constrained by the above considerations.
Acknowledgements
The authors would like to thank S. Brodsky, S. Chivukula, Y. Grossman, J. Ng, M. Peskin, A. Pomarol, M. Schmaltz, and J.Wells for discussions related to this work.
|
warning/0006/quant-ph0006105.html
|
ar5iv
|
text
|
# Entanglement molecules
## Abstract
We investigate the entanglement properties of multiparticle systems, concentrating on the case where the entanglement is robust against disposal of particles. Two qubits โbelonging to a multipartite systemโ are entangled in this sense iff their reduced density matrix is entangled. We introduce a family of multiqubit states, for which one can choose for any pair of qubits independently whether they should be entangled or not as well as the relative strength of the entanglement, thus providing the possibility to construct all kinds of โEntanglement moleculesโ. For some particular configurations, we also give the maximal amount of entanglement achievable.
Entanglement is at the heart of Quantum Information theory. In recent years, there has been an ongoing effort to characterice quantitatively and qualitatively entanglenglement. While for bipartite systems this problem is essentially solved, it remains still open for multipartite systems. In this case, there exist several possible approaches to identify different kinds of multipartite entanglement, and many interesting phenomena related to multipartite entanglement have been discovered .
Here we concentrate on bipartite aspects of multipartite entanglement, in particular on bipartite entanglement which is robust against disposal of particles. We consider $`N`$ spatially separated parties $`A_1,\mathrm{},A_N`$, each possessing a qubit. We first investigate the $`N`$โparty Greenberger-Horne-Zeilinger (GHZ)-state
$$|GHZ=\frac{1}{\sqrt{2}}(|0^N+|1^N),$$
(1)
which is considered to be a maximally entangled state (MES) of $`N`$ particles in several senses. For example, one can create a MES shared by any two of the parties with help of a (local) measurement performed by the remaining ones. Thus any two particles are potentially entangled, i.e. when allowing for assistance of the other parties, bipartite entanglement can be obtained from state $`|GHZ`$. However, it is essential that the remaining $`(N2)`$ parties perform measurements to assist the other two parties to share entanglement. If however โfor some reasonโ the information about only one of the particles, say $`A_N`$, is lost (or party $`A_N`$ decides not to cooperate with the remaining ones), the state of the remaining parties is only classically correlated and thus not entangled. In particular, the reduced density operator of any two parties is separable<sup>*</sup><sup>*</sup>*Given a $`N`$โpartite state $`\rho `$, the reduced density operator $`\rho _{12}`$ of party $`A_1`$ and $`A_2`$ is defined as $`\rho _{12}\mathrm{tr}_{3,\mathrm{},N}(\rho )`$. The operator $`\rho _{12}`$ is separable if it can be written as a convex combination of product states.. When considering the reduced density operator of two parties, we deal with the situation where the information about all remaining particles is not accessable (or the remaining parties are not willing to cooperate).
We thus say that two particles are entangled iff their reduced density operator is nonโseparable, i.e. the two particles share entanglement, independent what happens to the remaining particles. Such a definition is very suitable from a practical point of view, as there are certain multipartite scenarious where one is interested in entanglement properties of pairs of parties, which are independent of other parties. Note that in this sense, the state $`|GHZ`$ contains no (bipartite) entanglement at all.
But are there $`N`$โparticle states which are still entangled when tracing out any $`(N2)`$ particles, i.e. are there states where all particles are entangled with all other particles ? And if this is possible, what is the maximal amount of entanglement the remaining two parties can share ? In this paper, we will answer these questions and we will consider the more general setup where some parties are entangled, while some others are not. For example, for $`N=3`$ one may have that the reduced density operators $`\rho _{12}`$ and $`\rho _{13}`$ remain entangled, but $`\rho _{23}`$ is separable. We will show that one can have all possible configurations of this kind, i.e. there exist states where one can choose for each of the reduced density operators $`\rho _{kl}`$, $`k<l\{1,\mathrm{},N\}`$ independently whether it should be entangled or not. This allows to build general structures of $`N`$ particle states, which we call โEntanglement moleculesโ in spirit of the generalization of Wootters idea of an โEntangled chainโ , where one has a string of qubits, each qubit being entangled only with its nearest neighbours. We are considering more general setups, e.g. closed rings of particles where one only has nearest neighbour entanglement, (finite) strings with distance dependent entanglement (also 2nd and 3rd and so on neighbourhood entanglement), entanglement buckiballs (like the $`C_{60}`$ molecule), or more generally all possible setups of this kind one may imagine.
There are $`N(N1)/2`$ different bipartite reduced density operators $`\rho _{kl}`$ which may be either separable or not. If the reduced density operator $`\rho _{kl}`$ is nonโseparable, this automatically implies that a MES shared between parties $`A_k`$ and $`A_l`$ can be distilled (when allowing for several copies of the state), even without help of the remaining parties. This is due to the fact that for two qubit systems, inseparability is equivalent to distillability . In fact, the remaining parties can by no means prevent parties $`A_k`$ and $`A_l`$ to distill a MES. In a diagram, this will be visualized by a line between particles $`A_k`$ and $`A_l`$ which represents entanglement between the two parties in question (see Fig. 1). Each of the particles may be entangled with one (or more) of the remaining $`(N1)`$ particles. In particular one can have that any particle is entangled with all the remaining ones. Clearly, it is interesting to ask how strong these โbindingsโ (the entanglement between two particles) can be. Therefore, one has to quantify the entanglement of the bipartite reduced density operators $`\rho _{kl}`$. In this work, we choose as a measure of entanglement the concurrence $`๐`$ (for a definition of the concurrence see e.g. ). On one hand, we follow the lines suggested in , on the other we have that the entanglement of formation โthe amount of entanglement required to prepare a state $`\rho `$โ is monotonically increasing with the concurrence $`๐`$ The entanglement of formation is given by $`E_f(\rho )=h(\frac{1}{2}+\frac{1}{2}\sqrt{1๐^2})`$, where $`๐`$ is the concurrence and $`h`$ is the binary entropy function $`h(x)=x\mathrm{log}_2x(1x)\mathrm{log}_2(1x)`$. and thus the concurrence itself may be used to measure the strength of the bindings. For two special cases of particular interest, we will give the states with the maximal achieveable strength of the bindings.
Let us start by introducing a family of $`N`$ qubit states, which includes all possible configurations of โEntanglement moleculesโ. First we specify for each of the reduced density operators $`\rho _{kl}`$ whether it should be distillable or not, i.e. whether entanglement between the parties $`A_k`$ and $`A_l`$ can be distilled โ without help of the remaining parties โ or not. Let $`I=\{k_1l_1,\mathrm{},k_Ml_M\}`$ be the set of all those pairs where distillation should be possible, i.e. for $`klI`$, we have that $`\rho _{kl}`$ is distillable. We define the state
$$|\mathrm{\Psi }_{ij}|\mathrm{\Psi }^+_{ij}|0\mathrm{}0_{\mathrm{rest}},$$
(2)
that is the particles $`A_i`$ and $`A_j`$ are in a MES, namely $`|\mathrm{\Psi }^+=1/\sqrt{2}(|01+|10)`$, and the remaining particles are disentangled from each other and from $`A_iA_j`$. We now introduce a family of states which has the desired properties:
$$\rho _I=\frac{1}{M}\underset{klI}{}x_{kl}|\mathrm{\Psi }_{kl}\mathrm{\Psi }_{kl}|,$$
(3)
where $`M_{klI}x_{kl}`$ is a normalization factor. It is straightforward to calculate the reduced density operators $`\rho _{kl}`$. In the standard basis, one can check for $`klI`$ that $`\rho _{kl}`$ is of the form
$$\rho _{kl}=\frac{1}{M}\left(\begin{array}{cccc}a& 0& 0& 0\\ 0& b& x_{kl}/2& 0\\ 0& x_{kl}/2& c& 0\\ 0& 0& 0& 0\end{array}\right),$$
(4)
while for $`mnI`$ one obtains
$$\rho _{mn}=\frac{1}{M}\left(\begin{array}{cccc}\stackrel{~}{a}& 0& 0& 0\\ 0& \stackrel{~}{b}& 0& 0\\ 0& 0& \stackrel{~}{c}& 0\\ 0& 0& 0& 0\end{array}\right),$$
(5)
It is simple to calculate the concurrence in both cases:
$`๐_{kl}`$ $`=`$ $`{\displaystyle \frac{x_{kl}}{M}}\text{ iff }klI`$ (6)
$`๐_{mn}`$ $`=`$ $`0\text{ iff }mnI,`$ (7)
where we denote $`๐_{kl}๐(\rho _{kl})`$. We have that in $`\mathrm{I}\mathrm{C}^2\mathrm{I}\mathrm{C}^2`$ systems, nonzero concurrence automatically implies distillability of the state, while zero concurrence implies separability. Thus it is already clear that $`\rho _I`$ has the desired properties, i.e. entanglement between $`A_k`$ and $`A_l`$ can be distilled iff $`klI`$. We see that we can arbitrarily choose the relative strength of the bindings (measured by the concurrence) via the positive coefficients $`x_{kl}`$. It is now straightforward to explicitly construct the examples illustrated in Fig. 1: In (a) we have that $`x_{12}=x_{34}=x_{56}=2/9,x_{23}=x_{45}=x_{16}=1/9`$ (entanglement ring); In (b) $`x_{kl}=1/6`$ iff both $`k`$ and $`l`$ are even \[odd\] respectively and zero otherwise (all even \[odd\] particles are equally entangled) ; In (c) we have that $`x_{1l}=1/5`$, while all other $`x_{kl}=0`$ (one particle equally entangled with five other ones); Finally, in (d) $`x_{kl}=1/15`$ (all particles equally entangled). In a similar way, one can construct all examples mentioned in the introduction, such as strings with distance dependent entanglement or entanglement buckiballs.
Note that in this construction, we have that
$$\underset{i<k}{}๐_{ik}=1,$$
(8)
which shows that in a situation where all bindings have the same strength, the concurrence of the corresponding reduced density operators is determined by the total number of bindings, i.e. the number of elements in $`I`$. Hence, a state constructed in this way, where each particle is entangled with all others (i.e. all possible reduced density operators are distillable) and the strength of all bindings is equal has $`๐_{kl}=2/[N(N1)]`$. As we shall see next, this is however not the maximum value one can achieve for this particular configuration. In the following, we will analyze the two situations illustrated in Fig. 1(c) and (d), namely โ(i) all particles are equally entangled and (ii) one particle which is equally entangled with $`(N1)`$ othersโ and determine the maximum strength of the bindings for $`N=3`$.
(i) All particles pairwise entangled
We start with the case where all particles are equally entangled with each other (see also Fig. 1(d)). As shown in , the state
$$|W=1/\sqrt{3}(|001+|010+|100),$$
(9)
is the state of three qubits whose entanglement has the highest degree of endurance against loss of one of the three qubits. In particular, $`|W`$ maximizes the function
$$_{\mathrm{min}}(\psi )\mathrm{min}(๐_{12},๐_{13},๐_{23}),$$
(10)
and has $`๐_W๐_{12}=๐_{13}=๐_{23}=2/3`$. It follows that $`|W`$ is the 3โqubit state where all particles are equally entangled and the entanglement โ measured by the concurrence $`๐`$โ is maximal. Comparing $`|W`$ with $`\rho _I`$ of the form (3) with $`I=\{12,13,23\}`$, one immeadetly observes that $`๐_{\rho _I}=1/3`$, while $`๐_W=2/3`$.
More generally, let us consider the $`N`$โparty form $`|W_N`$ of the state $`|W`$, defined as
$$|W_N1/\sqrt{N}|N1,1,$$
(11)
where $`|N1,1`$ denotes the (unnormalized) totally symmetric state including $`N1`$ zeros and $`1`$ ones, e.g. $`|2,1=|001+|010+|100`$. As shown in , state $`|W_N`$ is a $`N`$ qubit state with all reduced density operators equal and the concurrences are given by $`๐_{kl}=2/N`$, which has to be compared to $`๐_{\rho _I}=2/[N(N1)]`$, $`\rho _I`$ being a state of the family (3) where all particles are equally entangled. It is however not known whether $`๐_{kl}=2/N,kl`$ is the maximal value achievable.
(ii) One particle entangled with $`N1`$ others
We consider now the case where one particle is equally entangled with $`(N1)`$ others and determine the maximal possible strength $`๐`$ of the bindings (see Fig. 1(c)). A related problem of this kind, namely the question of optimal entanglement splitting, i.e. the optimal way for a party $`B`$ to equally distribute its initial entanglement (shared with a party $`A`$) among several partners, was recently analyzed by Bruร in .
As shown in , we have for $`N=3`$ that $`๐_{12}^2+๐_{13}^21`$, i.e for $`๐_{12}=๐_{13}๐`$ we have that $`๐1/\sqrt{2}`$. This value is achieved by the state
$$|\psi =\frac{1}{\sqrt{2}}|100+\frac{1}{2}(|001+|010).$$
(12)
More generally, it was conjectured in that the inequality $`_{k=2}^N๐_{1k}^21`$ also holds. If we demand again that $`๐_{1k}=๐_{1l}๐\text{ }k,l>1`$, we find that $`๐1/\sqrt{N1}`$. This value is obtained by the state
$$|\psi =a|1|00\mathrm{}0+b|0|N2,1),$$
(13)
with $`a=1/\sqrt{2}`$ and $`b=1/\sqrt{2(N1)}`$, where $`|N2,1`$ is again an (unnormalized) totally symmetric state including $`(N2)`$ zeros and one 1. For a state of the form (3), where one particle is equally entangled with $`(N1)`$ others, one obtains $`๐_{1k}=1/(N1)`$.
For practical purposes, it may sometimes be useful to consider the fidelity of the reduced density operator (i.e. the maximal overlap of the reduced density operator with any MES) โwhich is generally not a proper measure of entanglementโ instead of the concurrence. The fidelity of the reduced density operator however indicates the achievable quality to perform certain quantum information tasks, e.g. teleportation . The fidelity of a density operator $`\rho `$ is defined as
$$F_\rho =\text{max}\mathrm{\Phi }|\rho |\mathrm{\Phi },$$
(14)
where the maximum is taken over all maximally entangled states $`|\mathrm{\Phi }`$. In the situation we consider, where one particle is equally entangled (i.e. $`F_\rho `$ is equal for all reduced density operators $`\rho _{1k}`$) with $`(N1)`$ others, one can derive a bound for the maximum fidelity $`F_\rho `$ of the reduced density operators $`\rho _{1k}`$ using results from optimal cloning . Given a density operator $`\rho `$ with a certain fidelity $`F_\rho `$, one can use $`\rho `$ to teleport the (unknown) state of a particle. As shown by the Horodecki , one can find a teleportation protocol which works equally well for all input states and has the maximum teleport fidelity $`F_t=(2F_\rho +1)/3`$. In our situation, we have that particle $`A_1`$ is entangled with $`(N1)`$ other particles. Thus, when performing a certain teleportation protocol , one obtains $`(N1)`$ (imperfect) clones of a state at the locations $`A_2,\mathrm{},A_N`$, where the quality of the clones is determined by the teleport fidelity $`F_t`$ of the corresponding reduced density operators. As shown by Werner , the maximum cloning fidelity of a $`1(N1)`$ cloner is given by $`F_c=[2(N1)+1]/[3(N1)]`$, from which follows that the teleport fidelity $`F_t`$ of the reduced density operators must fulfill $`F_tF_c`$, otherwise one could construct in this way a cloning machine which works better than the optimal one, which is clearly impossible. We thus have that
$$F_\rho \frac{1}{2}+\frac{1}{2(N1)}.$$
(15)
For the state $`|\psi `$ which maximizes the concurrence, we find that it does not obtain the maximal possible fidelity $`F_\rho `$. However, there exist states for which the maximal possible value of the fidelity is obtained. For example, a state of the form (13) with $`a=1/\sqrt{N(N1)}`$ and $`b=\sqrt{(N1)/N}`$ has the desired properties, as well as the โtelecloning stateโ introduced in . On the other hand, a state of the form (3), where one particle is equally entangled with $`(N1)`$ others has $`F_\rho =[2+N]/[4(N1)]`$, much smaller than the optimal value.
Note that the maximum value of the entanglement (measured by the concurrence) of a given qubit with its neighbours is not simply determined by the number of entangled neighbours, but also by the properties of the neighbours (i.e. the number of particles to which the neighbouring particles are entangled). This can be seen by noting that (i) in the case where one particle is equally entangled with two others (which are disentangled among themselves), the maximimal value for the concurrence is given by $`๐=1/\sqrt{2}`$ and (ii) in the case where three qubits are equally entangled, the maximum value is given by $`๐=2/3`$, which shows that the entanglement between systems $`A_2`$ and $`A_3`$ influence the maximum value of the entanglement between systems $`A_1A_2`$ and $`A_1A_3`$.
In summary, we provided a family of states which allows to construct all possible kinds of โEntanglement moleculesโ, i.e. one can choose for any pair of qubits independently whether a MES can be distilled without help of the remaining parties or not. In addition, the relative strength of the bindings (measured by the concurrence) can be adjusted arbitrarily. We investigated two particular configurations closer and provided states achieving the maximum value for the strength of the bindings.
I would like to thank G. Vidal and J. I. Cirac for discussions. This work was supported by the Austrian Science Foundation under the SFB โcontrol and measurement of coherent quantum systems (Project 11), the European Community under the TMR network ERBโFMRXโCT96โ0087 and project EQUIP (contract IST-1999-11053), the European Science Foundation, and the Institute for Quantum Information GmbH.
|
warning/0006/gr-qc0006016.html
|
ar5iv
|
text
|
# An Approximate Model of the Spacetime Foam
## I Introduction
The notion of a spacetime foam was introduced by Wheeler for the description of the possible complex structure of the spacetime on the Planck scale ($`L_{Pl}10^{33}cm`$). The exact mathematical description of this phenomenon is very difficult and even though there is a doubt: does the Feynman path integral in the gravity contain a topology change of the spacetime ? This question spring up as (according to the Morse theory) the singular points must arise by topology changes. In such points the time arrow is undefined that leads in difficulties at definition of the Lorentzian metric, curvature tensor and so on.
Here we propose an effective model of the spacetime foam in which a spinor field is introduced for an approximate description of the foam. For such model it is necessary the nonminimal interaction between spinor and electromagnetic fields (Pauli term). We will show that such interaction exists in the 5D Kaluza-Klein theory with a spinor field in such a way that the corresponding Maxwell equation is very similar to the electrodynamic in the continuous media.
In Ref. is presented a model of the wormhole in which a throat is a cloud of quantum wormholes (QWH) (see Fig.(1)).
To describe these QWHs we introduce a spinor field. In fact the spinor field is used for some approximate and effective description of QWHs as we are not able to do it by a direct way. Here we would like to offer the following model of the spacetime foam:
1. Each QWH is a solution of the 5D vacuum Einstein equations with $`G_{5t},G_{5\phi }0`$ components of the metric that leads to the appearance of electric and magnetic fields. In some approximation we can neglect of all linear sizes of QWH and obtain that Smolin calls as a โminimalistโ wormhole. For the 4D observer each mouth look as $`(\pm )`$ electric charge since this mouth entraps the force lines of the electric field (see Fig.(3)).
2. For the external observer each QWH is like to dipole and the spacetime with the foam seems as a dielectric by filled dipoles (see Fig.(2)).
3. The spacetime foam is described by a spinor field $`\psi `$ and the physical meaning of $`\psi `$ depends on an interaction term between spinor and electromagnetic fields that we shall discuss below.
4. An interaction between electromagnetic and spinor fields is nonminimal that allows us to interpret the Maxwell equations like to the electrodynamic in a continuous media.
## II Model of the individual quantum wormhole
The model of the individual QWH is presented on the Fig.(3). In fact this is some realization of the Wheeler idea about a wormhole entrapping electric force lines. In Ref. he wrote: โAlong with the fluctuations in the metric there occur fluctuations in the electromagnetic field. In consequence the typical multiply connected space $`\mathrm{}`$ has a net flux of electric lines of force passing through the โwormholeโ. These lines are trapped by the topology of the space. These lines give the appearance of a positive charge at one end of the wormhole and a negative charge at the otherโ.
The composite wormhole on the Fig.(3) consists from two Reissner-Nordstrรถm black holes and the 5D throat inserted between them . The 5D metric for this throat is
$$ds_{(5)}^2=R_0^2e^{2\psi (r)}\mathrm{\Delta }(r)\left(d\chi +\omega (r)dt+Q\mathrm{cos}\theta d\phi \right)^2+\frac{1}{\mathrm{\Delta }(r)}dt^2dr^2a(r)\left(d\theta ^2+\mathrm{sin}\theta ^2d\phi ^2\right),$$
(1)
where $`\chi `$ is the 5<sup>th</sup> extra coordinate; $`R_0>0`$ and $`Q`$ are some constants. And we assume that in some approximation such QWH of spacetime foam can be presented by this manner.
The 5D Einstein equations are
$`{\displaystyle \frac{\mathrm{\Delta }^{\prime \prime }}{\mathrm{\Delta }}}{\displaystyle \frac{\mathrm{\Delta }_{}^{}{}_{}{}^{2}}{\mathrm{\Delta }^2}}+{\displaystyle \frac{\mathrm{\Delta }^{}\psi ^{}}{\mathrm{\Delta }}}+{\displaystyle \frac{a^{}\mathrm{\Delta }^{}}{a\mathrm{\Delta }}}+R_0^2\omega _{}^{}{}_{}{}^{2}\mathrm{\Delta }^2e^{2\psi }`$ $`=`$ $`0,`$ (2)
$`\omega ^{\prime \prime }+\omega ^{}\left(2{\displaystyle \frac{\mathrm{\Delta }^{}}{\mathrm{\Delta }}}+3\psi ^{}+{\displaystyle \frac{a^{}}{a}}\right)`$ $`=`$ $`0,`$ (3)
$`{\displaystyle \frac{a^{\prime \prime }}{a}}+{\displaystyle \frac{a^{}\psi ^{}}{a}}{\displaystyle \frac{2}{a}}+{\displaystyle \frac{Q^2\mathrm{\Delta }e^{2\psi }}{a^2}}`$ $`=`$ $`0,`$ (4)
$`\psi ^{\prime \prime }+\psi _{}^{}{}_{}{}^{2}+{\displaystyle \frac{a^{}\psi ^{}}{a}}{\displaystyle \frac{Q^2\mathrm{\Delta }e^{2\psi }}{2a^2}}`$ $`=`$ $`0,`$ (5)
$`{\displaystyle \frac{\mathrm{\Delta }_{}^{}{}_{}{}^{2}}{\mathrm{\Delta }^2}}+2{\displaystyle \frac{\mathrm{\Delta }^{}\psi ^{}}{\mathrm{\Delta }}}4{\displaystyle \frac{a^{}\psi ^{}}{a}}+{\displaystyle \frac{4}{a}}{\displaystyle \frac{a_{}^{}{}_{}{}^{2}}{a^2}}R_0^2\omega _{}^{}{}_{}{}^{2}\mathrm{\Delta }^2e^{2\psi }{\displaystyle \frac{Q^2\mathrm{\Delta }e^{2\psi }}{a^2}}`$ $`=`$ $`0.`$ (6)
In Ref. it is shown that there is three type of solutions : the first type (wormhole-like solution) is presented on Fig.(3) with $`E>H`$ ($`E`$ and $`H`$ are Kaluza-Klein electric and magnetic fields), the second one is an infinite flux tube with $`E=H`$ and the third one is a singular solution (finite flux tube) with $`E<H`$. The definitions for $`E`$ and $`H`$ fields will be given later.
The longitudinal size $`l_0`$ of the WH-like solution depends on the relation between $`E`$ and $`H`$ : if $`H/E1`$ then $`l_0\mathrm{}`$. Let us define an approximate solution close to points $`r^2=r_0^2`$ (where $`ds^2(\pm r_0)=0`$). This solution we search in the form
$`\mathrm{\Delta }`$ $``$ $`\mathrm{\Delta }_1\left(r_0^2r^2\right),`$ (7)
$`\omega `$ $``$ $`{\displaystyle \frac{\omega _1}{r_0^2r^2}},`$ (8)
$`\psi `$ $``$ $`{\displaystyle \frac{\psi _3}{6}}\left(r_0^2r^2\right)^3.`$ (9)
The solution is
$`\mathrm{\Delta }_1`$ $`=`$ $`\pm {\displaystyle \frac{q}{2a_0r_0}},`$ (10)
$`\omega _1`$ $`=`$ $`{\displaystyle \frac{2a_0r_0}{q}},`$ (11)
$`\psi _3`$ $`=`$ $`\pm {\displaystyle \frac{qQ^2}{2a_0^3r_0^3}}`$ (12)
here $`a_0=a(r=\pm r_0)`$, $`q`$ is some constant. It is easy to show that at the hypersurfaces $`r=\pm r_0`$ : $`ds^2=0`$. On these hypersurfaces the change of the metric signature takes place : $`(+,,,,)`$ by $`|r|<r_0`$ and $`(,,,,+)`$ by $`|r|>r_0`$. Following to Bronnikov we call these two hypersurfaces as $`T`$horizons.
For the definition of a Kaluza-Klein electric field we consider Eq.(3)
$$\left[\left(\omega ^{}\mathrm{\Delta }^2e^{3\psi }\right)4\pi a\right]^{}=0$$
(13)
here $`4\pi a`$ is the area of $`S^2`$ sphere. Comparing with the Gauss law we see that Kaluza-Klein electric field can be defined as follows
$$E_{KK}=\omega ^{}\mathrm{\Delta }^2e^{3\psi }=\frac{q}{a}$$
(14)
here $`q`$ is an electric charge which is proportional to a flux of electric field. In this case the force lines of the electric field are uninterrupted and can be continued through the surfaces of matching the 5D WH-like solution and the Reissner-Nordstrรถm solution like to Fig.3. For the definition of a Kaluza-Klein magnetic field we write the following 5D Einstein equation
$$R_{\chi \phi }=\frac{e^\psi \sqrt{\mathrm{\Delta }}}{a^2\mathrm{sin}\theta }\frac{}{\theta }(Q)=0$$
(15)
and compare it with the ordinary 4D Maxwell equation
$$\frac{1}{\sqrt{g}}\frac{}{x^\mu }(\sqrt{g}F^{\phi \mu })=\frac{1}{\mathrm{sin}\theta }\frac{}{\theta }(\mathrm{sin}\theta F^{\theta \phi })=\frac{1}{\mathrm{sin}\theta }\frac{}{\theta }(\frac{H_r}{a})=0$$
(16)
here $`F^{\theta \phi }=\frac{ฯต^{\theta \phi i}}{\sqrt{\gamma }}H_i`$, $`\epsilon ^{ijk}`$ is the antisymmetrical tensor, $`\gamma `$ is the determinant of the 3D metric. The result is
$$H_r=\frac{Q\sqrt{\mathrm{\Delta }}e^\psi }{a}.$$
(17)
Immediately we see that $`H_r0`$ by $`r\pm r_0`$. The Einstein equations tell us that close to hypersurface $`r=\pm r_0`$ the Kaluza-Klein magnetic field can not have any influence on the gravity as the following term in Eqโs (4), (5) and (6) tends to zero
$$H_r^2=\frac{Q^2\mathrm{\Delta }e^{2\psi }}{a^2}0\text{by}r\pm r_0.$$
(18)
It means that the WH-like solutions near to these hypersurfaces are identical to the solution without the magnetic field. The external 4D observer sees that the force lines of magnetic field do not cross the event horizon for such composite WH. Another words each of QWH is like to moving electric charge but not a magnetic charge.
On these $`T`$horizons we should match:
* the flux of the 4D electric field (defined by the Maxwell equations) with the flux of the 5D electric field defined by $`R_{5t}=0`$ Kaluza-Klein equation.
* the area of the Reissner-Nordstrรถm event horizon with the area of the $`T`$hprizon.
It is necessary to note that both solutions (Reissner-Nordstrรถm black hole and 5D throat) have only two integration constants<sup>*</sup><sup>*</sup>*in fact, for the Reissner-Nordstrรถm black hole this leads to the โno hairโ theorem. and on the event horizon takes place an algebraic relation between these 4D and 5D integration constants. Another explanation of the fact that we use only two joining condition is the following (see Ref. for the more detailed explanations): in some sense on the event horizon holds a โholography principleโ. This means that in the presence of the event horizon the 4D and 5D Einstein equations lead to a reduction of the amount of initial data. For example the Einstein - Maxwell equations for the Reissner-Nordstrรถm metric
$`ds^2`$ $`=`$ $`\mathrm{\Delta }dt^2{\displaystyle \frac{dr^2}{\mathrm{\Delta }}}r^2\left(d\theta ^2+\mathrm{sin}^2d\phi ^2\right),`$ (19)
$`A_\mu `$ $`=`$ $`(\omega ,0,0,0))`$ (20)
(where $`A_\mu `$ is the electromagnetic potential, $`\kappa `$ is the gravitational constant) can be written as
$`{\displaystyle \frac{\mathrm{\Delta }^{}}{r}}+{\displaystyle \frac{1\mathrm{\Delta }}{r^2}}`$ $`=`$ $`{\displaystyle \frac{\kappa }{2}}\omega _{}^{}{}_{}{}^{2},`$ (21)
$`\omega ^{}`$ $`=`$ $`{\displaystyle \frac{q}{r^2}}.`$ (22)
For the Reissner - Nordstrรถm black hole the event horizon is defined by the condition $`\mathrm{\Delta }(r_g)=0`$, where $`r_g`$ is the radius of the event horizon. Hence in this case we see that on the event horizon
$$\mathrm{\Delta }_g^{}=\frac{1}{r_g}\frac{\kappa }{2}r_g\omega _{g}^{}{}_{}{}^{2},$$
(23)
here (g) means that the corresponding value is taken on the event horizon. Thus, Eq. (21), which is the Einstein equation, is the first-order differential equation in the whole spacetime $`(rr_g)`$. The condition (23) tells us that the derivative of the metric on the event horizon is expressed through the metric value on the event horizon. This is the same what we said above: the reduction of the amount of initial data takes place by such a way that we have only two integration constants (mass $`m`$ and charge $`e`$ for the Reissner-Nordstrรถm solution and $`q`$ and $`r_0`$ for the 5D throat).
## III Spacetime Foam and Spinor Fields
On the next approximation step we want to neglect with a cross section and longitudinal length of the 5D throat. In the result each QWH looks as an identification of two points, see Fig.4
Following to Smolin we introduce an operator $`\widehat{A}^{ab}(x^\mu ,y^\mu )`$ describing a quantum state in which the space with two points $`x`$ and $`y`$ fluctuates between two possibilities : points $`(x,y)`$ either are pasted together or not. In fact this operator describes an undeterminacy connected with the creation/annihilation of a wormholeThis is like to spin : $`z`$projection of spin can have two values $`\pm \mathrm{}/2`$.. Smolin calls such wormhole as a minimalist wormhole. The minimalist wormhole can be received from the above-mentioned composite wormhole if we neglect the linear sizes of 5D throat, i.e. shrink their to a point (see, Fig.4). We demand that this operator $`\widehat{A}^{ab}(x,y)`$ should have the following property
$$\widehat{A}^{ab}(x,y)=\theta ^a(x)\theta ^b(y)$$
(24)
$`a,b`$ are some indices which will be determine later. Of coarse for the definition of $`\theta ^a(x)`$ we should have some additional equation for this quantity. Smolinโs definition is
$`\widehat{A}^{ab}(x,y)`$ $`=`$ $`ฯต^{ab}\widehat{A}(x,y)=\theta ^a(x)\theta ^b(y),`$ (25)
$`\left(\widehat{A}(x,y)\right)^2`$ $`=`$ $`0`$ (26)
here $`a,b`$ are the spinor indices and $`\theta ^a(x)`$ is a spinor, $`ฯต^{12}=ฯต^{21}=1,ฯต^{11}=ฯต^{22}=0`$. We would like to say that it can be various definitions : a dynamical definition with field equations is given in section IV and the definition for which $`\theta ^a(x)=\theta ^a=const`$ will be given in section V.
## IV An effective model of the spacetime foam
In this case the quantity $`\theta ^a(x)`$ is a spinor field $`\psi ^\alpha (x)=\theta ^\alpha (x)`$ ($`a=\alpha `$ is the spinor index) and we determine $`\psi (x)`$ dynamically by means of some field equations for $`\psi (x)`$.
It is well known that gauge fields naturally appears in multidimensional gravities , , as some components of the multi-bein. But for the spinor field it is not the case. The spinor field in 4D and 5D spacetimes can have different interaction with gauge fields. For the 4D case the interaction term in Lagrangian is minimal $`(i\overline{\psi }A_\mu \gamma ^\mu \psi `$, $`\mu `$ is the 4D index) but for the second case it can be $`(i\overline{\psi }F_{AB}\gamma ^A\gamma ^B\psi )`$ or $`(iF_{AB}\overline{\psi }^A\psi ^B)`$ (here $`\psi ^A`$ is the Rarita-Schwinger spinor, $`A,B`$ are the 5D indices) or something like this.
We will consider the 5D Kaluza-Klein theory + torsion + spinor field with the 5D metric
$$ds^2=\left(d\chi +A_\mu dx^\mu \right)^2+g_{\mu \nu }dx^\mu dx^\nu $$
(27)
In this case (according to the initial interpretation of the Kaluza-Klein gravity with $`G_{55}=const`$) we have the electromagnetic potential $`A_\mu `$ and the 4D metric $`g_{\mu \nu }`$. The Lagrangian for this theory is
$``$ $`=`$ $`\sqrt{G}\{{\displaystyle \frac{1}{2k}}(R^{(5)}S_{ABC}S^{ABC})+`$ (29)
$`{\displaystyle \frac{\mathrm{}c}{2}}\{i\overline{\psi }[\gamma ^C(_C{\displaystyle \frac{1}{4}}\omega _{\overline{A}\overline{B}C}\gamma ^{[\overline{A}}\gamma ^{\overline{B}]}{\displaystyle \frac{1}{4}}S_{\overline{A}\overline{B}C}\gamma ^{[\overline{A}}\gamma ^{\overline{B}]}){\displaystyle \frac{mc}{i\mathrm{}}}]\psi +h.c.\}\}`$
where $`G`$ is the determinant of the 5D metric, $`R^{(5)}`$ is the 5D scalar curvature, $`S_{ABC}`$ is the antisymmetrical torsion tensor, $`A,B,C`$ are the 5D world indexes, $`\overline{A},\overline{B},\overline{C}`$ are the 5-bein indexes, $`\gamma ^B=h_{\overline{A}}^B\gamma ^{\overline{A}}`$, $`h_{\overline{A}}^B`$ is the 5-bein, $`\gamma ^{\overline{A}}`$ are the 5D $`\gamma `$ matrixes with usual definitions $`\gamma ^{\overline{A}}\gamma ^{\overline{B}}+\gamma ^{\overline{B}}\gamma ^{\overline{A}}=2\eta ^{\overline{A}\overline{B}}`$, $`\eta ^{\overline{A}\overline{B}}=(+,,,,)`$ is the signature of the 5D metric; $`[]`$ means the antisymmetrization, $`\mathrm{}`$, $`c`$ and $`m`$ are the usual constants. The most important for us is the choice of a spinor $`\psi `$ which will approximately describe the spacetime foam as it was mentioned in the section III : i.e. $`\psi ^\alpha (x)=\theta ^\alpha (x)`$. It is very important to note that in the context of this section we have some dynamical equations for the $`\theta ^\alpha (x)=\psi ^\alpha (x)`$ (it is convenient to use the usual designation $`\psi `$ for the fermion field). We should note that all physical fields in Lagrangian (29) must be quantum operators but on the first approximation step we change their by classical fields. After dimensional reduction we have
$``$ $`=`$ $`\sqrt{g}\{{\displaystyle \frac{1}{2k}}(R+{\displaystyle \frac{1}{4}}F_{\alpha \beta }F^{\alpha \beta })+`$ (31)
$`{\displaystyle \frac{\mathrm{}c}{2}}[i\overline{\psi }(\gamma ^\mu \stackrel{~}{}_\mu {\displaystyle \frac{1}{8}}F_{\overline{\alpha }\overline{\beta }}\gamma ^{\overline{5}}\gamma ^{[\overline{\alpha }}\gamma ^{\overline{\beta }]}{\displaystyle \frac{1}{4}}l_{Pl}^2\left(\gamma ^{[\overline{A}}\gamma ^{\overline{B}}\gamma ^{\overline{C}]}\right)\left(i\overline{\psi }\gamma _{[\overline{A}}\gamma _{\overline{B}}\gamma _{\overline{C}]}\psi \right){\displaystyle \frac{mc}{i\mathrm{}}})\psi +h.c.]\}`$
$`S^{\overline{A}\overline{B}\overline{C}}`$ $`=`$ $`2l_{Pl}^2\left(i\overline{\psi }\gamma ^{[\overline{A}}\gamma ^{\overline{B}}\gamma ^{\overline{C}]}\psi \right)`$ (32)
where $`g`$ is the determinant of the 4D metric, $`R`$ is the 4D scalar curvature, $`F_{\alpha \beta }=_\alpha A_\beta _\beta A_\alpha `$ is the Maxwell tensor, $`A_\mu =h_\mu ^{\overline{5}}`$ is the electromagnetic potential, $`\alpha ,\beta ,\mu `$ are the 4D world indexes, $`\overline{\alpha },\overline{\beta },\overline{\mu }`$ are the 4D vier-bein indexes, $`h_\nu ^{\overline{\mu }}`$ is the vier-bein, $`\gamma ^{\overline{\mu }}`$ are the 4D $`\gamma `$ matrixes with usual definitions $`\gamma ^{\overline{\mu }}\gamma ^{\overline{\nu }}+\gamma ^{\overline{\nu }}\gamma ^{\overline{\mu }}=2\eta ^{\overline{\mu }\overline{\nu }}`$, $`\eta ^{\overline{\mu }\overline{\nu }}=(+,,,)`$ is the signature of the 4D metric. Varying with respect to $`g_{\mu \nu }`$, $`\overline{\psi }`$ and $`A_\mu `$ leads to the following equations
$`R_{\mu \nu }{\displaystyle \frac{1}{2}}g_{\mu \nu }R={\displaystyle \frac{1}{2}}\left(F_{\mu \alpha }F_\nu ^\alpha +{\displaystyle \frac{1}{4}}g_{\mu \nu }F_{\alpha \beta }F^{\alpha \beta }\right)`$ $`+`$ $`4l_{Pl}^2[(i\overline{\psi }\gamma _\mu \stackrel{~}{}_\nu \psi +i\overline{\psi }\gamma _\nu \stackrel{~}{}_\mu \psi )+h.c.]`$ (33)
$`2l_{Pl}^2\left[F_{\mu \alpha }\left(i\overline{\psi }\gamma ^{\overline{5}}\gamma _{[\nu }\gamma ^{\alpha ]}\psi \right)+F_{\nu \alpha }\left(i\overline{\psi }\gamma ^{\overline{5}}\gamma _{[\mu }\gamma ^{\alpha ]}\psi \right)\right]`$ $``$ $`2g_{\mu \nu }l_{Pl}^4\left(i\overline{\psi }\gamma ^{[\overline{A}}\gamma ^{\overline{B}}\gamma ^{\overline{C}]}\psi \right)\left(i\overline{\psi }\gamma _{[\overline{A}}\gamma _{\overline{B}}\gamma _{\overline{C}]}\psi \right),`$ (34)
$`D_\nu H^{\mu \nu }=0,H^{\mu \nu }=F^{\mu \nu }+\stackrel{~}{F}^{\mu \nu };\stackrel{~}{F}^{\mu \nu }`$ $`=`$ $`4l_{Pl}^2\left(i\overline{\psi }\gamma ^{\overline{5}}\gamma ^{[\mu }\gamma ^{\nu ]}\psi \right)=4l_{Pl}^2E^{\mu \nu \alpha \beta }\left(i\overline{\psi }\gamma _{[\alpha }\gamma _{\beta ]}\psi \right),`$ (35)
$`i\gamma ^\mu \stackrel{~}{}_\mu \psi {\displaystyle \frac{1}{8}}F_{\overline{\alpha }\overline{\beta }}\left(i\gamma ^{\overline{5}}\gamma ^{[\overline{\alpha }}\gamma ^{\overline{\beta }]}\psi \right)`$ $``$ $`{\displaystyle \frac{1}{2}}l_{Pl}^2\left(i\gamma ^{[\overline{A}}\gamma ^{\overline{B}}\gamma ^{\overline{C}]}\psi \right)\left(i\overline{\psi }\gamma _{[\overline{A}}\gamma _{\overline{B}}\gamma _{\overline{C}]}\psi \right)=0,`$ (36)
$`\stackrel{~}{}_\mu `$ $`=`$ $`_\mu {\displaystyle \frac{1}{4}}\omega _{\overline{a}\overline{b}\mu }\gamma ^{[\overline{a}}\gamma ^{\overline{b}]}`$ (37)
where $`\stackrel{~}{}_\mu `$ is the 4D covariant derivative of the spinor field without torsion, $`\omega _{\overline{a}\overline{b}\mu }`$ is the 4D Ricci coefficients without torsion, $`E^{\mu \nu \alpha \beta }`$ is the 4D absolutely antisymmetric tensor. The most interesting for us is the Maxwell equation (35) which permits us to discuss the physical meaning of the spinor field. We would like to show that this equation in the given form is similar to the electrodynamic in the continuous media. Let we remind that for the electrodynamic in the continuous media two tensors $`\overline{F}^{\mu \nu }`$ and $`\overline{H}^{\mu \nu }`$ are introduced for which we have the following equations system (in the Minkowski spacetime)
$`\overline{F}_{\alpha \beta ,\gamma }+\overline{F}_{\gamma \alpha ,\beta }+\overline{F}_{\beta \gamma ,\alpha }`$ $`=`$ $`0,`$ (38)
$`\overline{H}_{,\beta }^{\alpha \beta }`$ $`=`$ $`0`$ (39)
and the following relations between these tensors
$`\overline{H}_{\alpha \beta }u^\beta `$ $`=`$ $`\epsilon \overline{F}_{\alpha \beta }u^\beta ,`$ (40)
$`\overline{F}_{\alpha \beta }u_\gamma +\overline{F}_{\gamma \alpha }u_\beta +\overline{F}_{\beta \gamma }u_\alpha `$ $`=`$ $`\mu \left(\overline{H}_{\alpha \beta }u_\gamma +\overline{H}_{\gamma \alpha }u_\beta +\overline{H}_{\beta \gamma }u_\alpha \right)`$ (41)
where $`\epsilon `$ and $`\mu `$ are the dielectric and magnetic permeability respectively, $`u^\alpha `$ is the 4-vector of the matter. For the rest media and in the 3D designation we have
$`\epsilon \overline{E}_i`$ $`=`$ $`\overline{E}_i+4\pi \overline{P}_i=\overline{D}_i,\text{where}\overline{E}_i=\overline{F}_{0i},\overline{D}_{0i}=\overline{H}_{0i},`$ (42)
$`\mu \overline{H}_i`$ $`=`$ $`\overline{H}_i+4\pi \overline{M}_i=\overline{B}_i,\text{where}\overline{B}_i=ฯต_{ijk}\overline{F}^{jk},\overline{H}_i=ฯต_{ijk}\overline{H}^{jk},`$ (43)
where $`P_i`$ is the dielectric polarization and $`M_i`$ is the magnetization vectors, $`ฯต_{ijk}`$ is the 3D absolutely antisymmetric tensor. Let us rewrite the definition in Eq.(35) in the following form
$`E_i+\stackrel{~}{E}_i`$ $`=`$ $`D_i\text{where }E_i=F_{0i},\stackrel{~}{E}_i=\stackrel{~}{F}_{0i},D_i=H_{0i}`$ (44)
$`B_i+\stackrel{~}{B}_i`$ $`=`$ $`H_i\text{where}B_i=ฯต_{ijk}F^{jk},\stackrel{~}{B}_i=ฯต_{ijk}\stackrel{~}{F}^{jk},H_i=ฯต_{ijk}H^{jk}.`$ (45)
Comparing these definitions with (42), (43) immediately we see that the following notations can be introduced
$$\stackrel{~}{E}_i=4l_{Pl}^2ฯต_{ijk}\left(i\overline{\psi }\gamma ^{[j}\gamma ^{k]}\psi \right)$$
(46)
is the polarization vector of the spacetime foam and
$$\stackrel{~}{B}_i=4l_{Pl}^2ฯต_{ijk}\left(i\overline{\psi }\gamma ^{\overline{5}}\gamma ^{[j}\gamma ^{k]}\psi \right)$$
(47)
is the magnetization vector of the spacetime foam.
The physical reason for this is : each QWH is like to a moving dipole (see Fig.(2)) which produces microscopical electric and magnetic fields.
## V Supergravity as a possible model of the spacetime foam
Now we would like to consider the another possible interpretation of $`\theta ^a(x)`$. In this case (in contrast to section IV) we determine the quantity $`\theta ^a(x)`$ by a non-dynamic way : introducing some algebraic equation (48) for $`\theta `$. Let the operator $`\widehat{A}^{ab}(x,y)`$ introduced in the section III satisfies to the following equation
$$\widehat{A}^{ab}(x,y)\widehat{A}_{ab}(x,y)=0.$$
(48)
The solution of Eq.(48) we search in the form
$$\widehat{A}^{ab}(x,y)=\theta ^a(x)\theta ^b(y).$$
(49)
After substitution in Eq.(48) we have
$$\theta ^a(x)\theta ^b(y)\theta _a(x)\theta _b(y)=0.$$
(50)
Like to section III we have two simplest solution. The first solution is
$`a`$ $`=`$ $`\alpha ,b=\beta ,(\alpha ,\beta =1,2),`$ (51)
$`\widehat{A}^{\alpha \beta }(x,y)`$ $`=`$ $`\theta ^\alpha (x)\theta ^\beta (y).`$ (52)
$`\theta ^\alpha `$ is an undotted spinor of $`(\frac{1}{2},0)`$ representation of Sl(2,C) group.
$`\theta ^\alpha (x)`$ $`=`$ $`\theta ^\alpha (y)=\theta ^\alpha =const,`$ (53)
$`\theta ^\alpha \theta ^\beta `$ $`=`$ $`\theta ^\beta \theta ^\alpha ;\alpha \beta ,`$ (54)
$`\left(\theta ^\alpha \right)^2`$ $`=`$ $`0.`$ (55)
The second solution is similar : $`\alpha \dot{\alpha }`$ and $`\beta \dot{\beta }`$
$`a`$ $`=`$ $`\dot{\alpha },b=\dot{\beta },(\dot{\alpha },\dot{\beta }=1,2),`$ (56)
$`\widehat{A}^{\dot{\alpha }\dot{\beta }}(x,y)`$ $`=`$ $`\overline{\theta }^{\dot{\alpha }}(x)\overline{\theta }^{\dot{\beta }}(y),`$ (57)
$`\overline{\theta }^{\dot{\alpha }}(x)`$ $`=`$ $`\overline{\theta }^{\dot{\alpha }}(y)=\overline{\theta }^{\dot{\alpha }}=const,`$ (58)
$`\overline{\theta }^{\dot{\alpha }}\overline{\theta }^{\dot{\beta }}`$ $`=`$ $`\overline{\theta }^{\dot{\beta }}\overline{\theta }^{\dot{\alpha }};\dot{\alpha }\dot{\beta },`$ (59)
$`\left(\overline{\theta }^{\dot{\alpha }}\right)^2`$ $`=`$ $`0.`$ (60)
In this case $`\overline{\theta }^{\dot{\alpha }}`$ is a dotted spinor of $`(0,\frac{1}{2})`$ representation. Such two-valuedness compels us to introduce both possibilities : $`\theta =\{\theta ^\alpha ,\overline{\theta }^{\dot{\alpha }}\}`$.
Like to Smolin we would like to introduce an infinitesimal operator $`\delta \widehat{B}(x^\mu z^\mu =x^\mu +\delta x^\mu )`$ of a displacement of the wormhole mouth
$`\delta \widehat{B}(x^\mu `$ $``$ $`z^\mu )\widehat{A}^{\gamma \delta }(y^\mu ,x^\mu )=\widehat{A}^{\gamma \delta }(y^\mu ,z^\mu ),`$ (61)
$`\widehat{A}^{\gamma \delta }(y^\mu ,x^\mu )`$ $`=`$ $`\theta ^\gamma \theta ^\delta ,`$ (62)
$`\widehat{A}^{\gamma \delta }(y^\mu ,z^\mu )`$ $`=`$ $`\theta _{}^{}{}_{}{}^{\gamma }\theta _{}^{}{}_{}{}^{\delta }=\left(\theta ^\gamma +\epsilon ^\gamma \right)\left(\theta ^\delta +\epsilon ^\delta \right)\theta ^\gamma \theta ^\delta +\epsilon ^\gamma \theta ^\delta \epsilon ^\delta \theta ^\gamma `$ (63)
here $`\epsilon ^\alpha `$ is an infinitesimal Grassmannian number. Therefore we have the following equation for the definition of $`\delta \widehat{B}(x^\mu z^\mu )`$ operator
$$\delta \widehat{B}\left(x^\mu z^\mu \right)\theta ^\gamma \theta ^\delta =\theta ^\gamma \theta ^\delta +\epsilon ^\gamma \theta ^\delta \epsilon ^\delta \theta ^\gamma .$$
(64)
This equation has the following solution
$$\delta \widehat{B}\left(x^\mu z^\mu \right)=1+\epsilon ^\alpha \frac{}{\theta ^\alpha }i\epsilon ^\alpha \sigma _{\alpha \dot{\beta }}^\mu \overline{\theta }^{\dot{\beta }}_\mu \overline{\epsilon }^{\dot{\alpha }}\frac{}{\overline{\theta }^{\dot{\alpha }}}+i\theta ^\alpha \sigma _{\alpha \dot{\beta }}^\mu \overline{\epsilon }^{\dot{\beta }}_\mu .$$
(65)
This allows us to say that $`\theta =\{\theta ^\alpha ,\theta ^{\dot{\alpha }}\}`$ are the Grassmanian numbers which we should use as some additional coordinates for the description of the spacetime foam. In this approach the superspace gravity with the anticommuting coordinates $`\theta `$ describes in some approximation the spacetime foam.
## VI Conclusions
Thus, here we have proposed the approximate model for the description of the spacetime foam. This model is based on the assumption that the whole spacetime is 5 dimensional but $`G_{55}`$ is the dynamical variable only in the QWHs. In this case 5D gravity has the solution which we have used as a model of the individual quantum wormhole. In the approximation when the 5D throat of each QWH is contracted to a point the spacetime foam can be approximately described by a spinor field or Grassmanian anticommuting coordinates on the superspace.
Such model leads to the very interesting experimental consequences. We see that the spacetime foam has 5D structure and it connected with the electric field. This observation allows us to presuppose that the very strong electric field can open a door into 5 dimension! The question is: as is great should be this field ? The electric field $`E_i`$ in the CGSE units and $`e_i`$ in the โgeometrizedโ units can be connected by formula
$`e_i`$ $`=`$ $`{\displaystyle \frac{G^{1/2}}{c^2}}E_i=\left(2.874\times 10^{25}cm^1/gauss\right)E_i,`$ (66)
$`\left[e_i\right]`$ $`=`$ $`cm^1,\left[E_i\right]=V/cm`$ (67)
As we see the value of $`e_i`$ is defined by some characteristic length $`l_0`$. It is possible that $`l_0`$ is a length of the $`5^{th}`$ dimension. If $`l_0=l_{Pl}`$ then $`E_i10^{57}V/cm`$ and this field strength is in the Planck region, and is will beyond experimental capabilities to create. But if $`l_0`$ has a different value it can lead to much more realistic scenario for the experimental capability to open door into $`5^{th}`$ dimension.
Another conclusion of this work is that : the supergravity theories can be considered as approximative and effective models of the spacetime foam.
## VII Acknowledgment
I am grateful for financial support from the Georg Forster Research Fellowship of the Alexander von Humboldt Foundation and H.-J. Schmidt for an invitation to Potsdam University.
|
warning/0006/nlin0006032.html
|
ar5iv
|
text
|
# I Introduction
## I Introduction
Supersymmetric integrable systems constitute a very interesting subject and as a consequence a number of well known integrable equations have been generalized into supersymmetric (SUSY) context. We mention the SUSY versions of sine-Gordon , , KP-hierarchy , KdV ,, Boussinesq etc. We also point out that there are many generalizations related to the number $`๐ฉ`$ of fermionic independent variables. In this paper we are dealing with the $`๐ฉ=1`$ superspace.
So far, many of the tools used in standard theory have been extended to this framework, such as Bรคcklund transformations , prolongation theory, hamiltonian formalism , grasmmannian description , $`\tau `$ functions , Darboux transformations . The physical interest in the study of these systems have been launched by the seminal paper of Alvarez-Gaume et. al about the partition function and super-Virasoro constraints of 2D quantum supergravity. Although the $`\tau `$ function theory in the context of SUSY pseudodifferential operators was given for the SUSY KP-hierarchy , the bilinear formalism for SUSY equations was very little investigated. We mention here the algebraic approach using the representation theory of affine Lie super-algebras in the papers of Kac and van der Leur , Kac and Medina the super-conformal field theoretic approach of LeClair . Anyway in these articles the bilinear hierarchies are not related to the SUSY hierarchies of nonlinear equations.
This paper which is an extended version of we consider a direct approach to SUSY equations in a $`๐ฉ=1`$ superspace rather than hierarchies namely extending the gauge-invariance principle of $`\tau `$ functions for classical Hirota operators. Our result generalize the Grammaticos-Ramani-Hietarinta theorem, to SUSY case and we find $`๐ฉ=1`$ SUSY Hirota bilinear operators. With these operators one can obtain SUSY-bilinear forms for SUSY KdV equation of Mathieu and also it allows bilinear forms for certain SUSY extensions of Sawada-Kotera-Ramani , Hirota-Satsuma , KdV-B , Burgers, mKdV and $`๐ฉ=2`$ -superspace SUSY sine-Gordon equations . Also the gauge-invariance principle allows to study the SUSY multisoliton solutions as exponentials of linears. A very interesting fact which happens is that SUSY KdV equation of Mathieu does not have 3 supersoliton solution for arbitrary choice of solitary waves although it possesses Lax pair . Only for special combination of parameters the equation admits N soliton solutions. Although this seems to be a quite strange paradox, Liu and Manas , found also super-soliton solutions for SUSY KdV equation in terms of pfaffians only for certain wave parameters. This fact shows that Hirota integrability and Lax integrability are different in the SUSY context.
The paper is organized as follows. In section II the standard bilinear formalism is briefly discussed. In section III supersymmetric versions for nonlinear evolution equations are presented and in section IV we introduce the super-bilinear formalism. In the last section we shall present the bilinear form for SUSY KdV-type equations, super-soliton solutions and several comments about extension to $`๐ฉ=2`$ SUSY sine-Gordon equation.
## II Standard bilinear formalism
The Hirota bilinear operators were introduced as an antisymmetric extension of the usual derivative , because of their usefulness for the computation of multisoliton solution of nonlinear evolution equations. The bilinear operator $`๐_x=_{x_1}_{x_2},`$ acts on a pair of functions (the so-called โdot productโ) antisymmetrically:
$$๐_xfg=(_{x_1}_{x_2})f(x_1)f(x_2)|_{x_1=x_2=x}=f^{}gfg^{}.$$
(1)
The Hirota bilinear formalism has been instrumental in the derivation of the multisoliton solutions of (integrable) nonlinear evolution equations. The first step in the application is a dependent variable transformation which converts the nonlinear equation into a quadratic form. This quadratic form turns out to have the same structure as the dispersion relation of the linearized nonlinear equation, although there is no deep reason for that. This is best understood if we consider an example. Starting from paradigmatic KdV equation
$$u_t+6uu_x+u_{xxx}=0,$$
(2)
we introduce the substitution $`u=2_x^2\mathrm{log}F`$ and obtain after one integration:
$$F_{xt}FF_xF_t+F_{xxxx}F4F_{xxx}F_x+3F_{xx}^2=0,$$
(3)
which can be written in the following condensed form:
$$(๐_x๐_t+๐_x^4)FF=0.$$
(4)
The power of the bilinear formalism lies in the fact that for multisoliton solution $`F`$โs are polynomials of exponentials. Moreover it displays also the interaction (phase-shifts) between solitons. In the case of KdV equation the multisoliton solution has the following form:
$$F=\underset{\mu =0,1}{}\mathrm{exp}(\underset{i=1}{\overset{N}{}}\mu _i\eta _i+\underset{i<j}{}A_{ij}\mu _i\mu _j),$$
(5)
where $`\eta _i=k_ixk_i^3t+\eta _i^{(0)}`$ and $`expA_{ij}=(\frac{k_ik_j}{k_i+k_j})^2`$ which is the phase-shift from the interaction of the soliton $`\mathrm{"}i\mathrm{"}`$ with the soliton $`\mathrm{"}j\mathrm{"}`$.
This picture can be generalized to any bilinear equation of the form
$$P(๐_\stackrel{}{x})FF=0$$
(6)
where $`P`$ is any polynomial and $`\stackrel{}{x}=(t,x,y,\mathrm{})`$. The 1 soliton solution is $`F=1+e^\eta `$, where $`\eta =\stackrel{}{k}\stackrel{}{x}+const.`$ This solution holds if
$$P(\stackrel{}{k})=0.$$
This is a condition on the parameters $`\stackrel{}{k}`$ of $`\eta `$ and is called the dispersion relation. If the parameter space is n-dimensional then the above equation defines an $`n1`$ dimensional submanifold called dispersion manifold.
Hirota ansatz for 2 soliton solution is
$$F=1+e^{\eta _1}+e^{\eta _2}+A_{12}e^{\eta _1+\eta _2}$$
where $`\eta `$โs are defined before and $`A_{12}`$ are function of $`\stackrel{}{k}_1`$ and $`\stackrel{}{k}_2`$ each one giving the coordinates of some point in the dispersion manifold. Substituting the ansatz in the equation (6) and taking into account the dispersion relation we find
$$A_{12}P(\stackrel{}{k}_1+\stackrel{}{k}_2)+P(\stackrel{}{k}_1\stackrel{}{k}_2)=0.$$
(7)
Generally speaking, the great majority of bilinear equations possess 2 soliton solution but only integrable equations possess 3 or more soliton solutions. The form of the N soliton solution is,
$$F=\underset{\mu =0,1}{}\mathrm{exp}(\underset{i=1}{\overset{N}{}}\mu _i\eta _i+\underset{i<j}{}A_{ij}\mu _i\mu _j)$$
More about the bilinear equations are in
A very important observation (which motivated the present paper) is the relation of the physical field $`u=2_x^2\mathrm{log}F`$ of KdV equation with the Hirota function $`F`$: the gauge-transformation $`Fe^{px+\omega t}F`$ leaves $`u`$ invariant. This is a general property of all bilinear equations. Moreover, one can define the Hirota operators using the requirement of gauge-invariance. Letโs introduce a general bilinear expression,
$$A_N(f,g)=\underset{i=0}{\overset{N}{}}c_i(_x^{Ni}f)(_x^ig)$$
(8)
and ask to be invariant under the gauge-trasformation:
$$A_N(e^\theta f,e^\theta g)=e^{2\theta }A_N(f,g)\theta =kx+\omega t+\mathrm{}(linears).$$
(9)
Then we have the following,
Theorem: $`A_N(f,g)`$ is gauge-invariant if and only if $`A_N(f,g)=๐_x^Nfg`$ i.e.
$$c_i=c_0(1)^i\left(\begin{array}{c}N\\ i\end{array}\right)$$
and $`c_0`$ is a constant and the brakets represent binomial coefficient.
We must point out that in the whole paper the natural number N (which will denote number of solitons, number of terms, exponents etc.) is different from $`๐ฉ`$ which is related to supersymmetry or superspace.
## III Supersymmetry
The supersymmetric extension of a nonlinear evolution equation (KdV for instance) refers to a system of coupled equations for a bosonic $`u(t,x)`$ and a fermionic field $`\xi (t,x)`$ which reduces to the initial equation in the limit where the fermionic field is zero (bosonic limit). In the classical context, a fermionic field is described by an anticommuting function with values in an infinitely generated Grassmann algebra. However, supersymmetry is not just a coupling of a bosonic field to a fermionic field. It implies a transformation (supersymmetry invariance) relating these two fields which leaves the system invariant. In order to have a mathematical formulation of these concepts we have to extend the classical space $`(x,t)`$ to a larger space (superspace) $`(t,x,\theta )`$ where $`\theta `$ is a Grassmann variable and also to extend the pair of fields $`(u,\xi )`$ to a larger fermionic or bosonic superfield $`\mathrm{\Phi }(t,x,\theta ).`$ In order to have nontrivial extension for KdV we choose $`\mathrm{\Phi }`$ to be fermionic, having the expansion
$$\mathrm{\Phi }(t,x,\theta )=\xi (t,x)+\theta u(t,x).$$
(10)
The $`๐ฉ=1`$ SUSY means that we have only one Grassmann variable $`\theta `$ and we consider only space supersymmetry invariance namely $`xx\lambda \theta `$ and $`\theta \theta +\lambda `$ ($`\lambda `$ is an anticommuting parameter). This transformation is generated by the operator $`Q=_\theta \theta _x,`$ which anticommutes with the covariant derivative $`D=_\theta +\theta _x`$ (Notice also that $`D^2=_x`$). Expressions written in terms of the covariant derivative and the superfield $`\mathrm{\Phi }`$ are manifestly supersymmetric invariant. Using the superspace formalism one can construct different supersymmetric extension of nonlinear equations. Thus the integrable (in the sense of Lax pair) variant of $`๐ฉ=1`$ SUSY KdV is
$$\mathrm{\Phi }_t+D^6\mathrm{\Phi }+3D^2(\mathrm{\Phi }D\mathrm{\Phi })=0,$$
(11)
which on the components has the form
$`u_t`$ $`=`$ $`u_{xxx}6uu_x+3\xi \xi _{xx}`$ (12)
$`\xi _t`$ $`=`$ $`\xi _{xxx}3\xi _xu3\xi u_x.`$ (13)
Another integrable variant of SUSY KdV equation which is very important in applications to supersymmetric matrix models is SUSY KdV-B equation , namely
$$\mathrm{\Phi }_t+D^6\mathrm{\Phi }+6D^2\mathrm{\Phi }D\mathrm{\Phi }=0,$$
(14)
leads to a somewhat trivial system in which the fermionic fields decouple from the bosonic equation which reduces then to the usual KdV.
We shall discuss also the following supersymmetric equations, although we do not know if it is completely integrable in the sense of Lax pair ($`\mathrm{\Phi }`$ is also a fermionic superfield).
* $`๐ฉ=1`$ SUSY Sawada-Kotera-Ramani,
$$\mathrm{\Phi }_t+D^{10}\mathrm{\Phi }+D^2(10D\mathrm{\Phi }D^4\mathrm{\Phi }+5D^5\mathrm{\Phi }\mathrm{\Phi }+15(D\mathrm{\Phi })^2\mathrm{\Phi })=0.$$
(15)
* $`๐ฉ=1`$ SUSY Hirota-Satsuma (shallow water wave)
$$D^4\mathrm{\Phi }_t+\mathrm{\Phi }_tD^3\mathrm{\Phi }+2D^2\mathrm{\Phi }D\mathrm{\Phi }_tD^2\mathrm{\Phi }\mathrm{\Phi }_t=0$$
(16)
* $`๐ฉ=1`$ SUSY Burgers
$$\mathrm{\Phi }_t+\mathrm{\Phi }D\mathrm{\Phi }_x+\mathrm{\Phi }_{xx}=0$$
(17)
A very important equation from the physical consideration is the SUSY sine-Gordon. We are going to consider the version studied by Kulish and Tsyplyaev . There are other integrable versions of SUSY sine-Gordon emerged from algebraic procedures . In this case one needs two Grassmann variables $`\theta _\alpha `$ with $`\alpha =1,2`$ and the supersymmetry transformation is
$$x^{}_{}{}^{}\mu =x^\mu i\overline{\lambda }\gamma ^\mu \theta ,\theta _\alpha ^{^{}}=\theta _\alpha +\lambda _\alpha ,\mu =1,2.$$
Here, $`\lambda _\alpha `$ is the anticommuting spinor parameter of the transformation and $`\overline{\lambda }=(\lambda ^1,\lambda ^2)`$, $`\lambda ^\alpha =\lambda _\beta (i\sigma _2)^{\beta \alpha }`$, $`\gamma ^0=i\sigma _2`$, $`\gamma ^1=\sigma _1`$, $`\gamma ^5=\gamma ^0\gamma ^1=\sigma _3`$. We use the metric $`g^{\mu \nu }=diag(1,1)`$ and $`\sigma _i`$ are the Pauli matrices. The superfield has the following expansion:
$$\mathrm{\Phi }(x^\mu ,\theta _\alpha )=\varphi (x^\mu )+i\overline{\theta }\psi (x^\mu )+\frac{i}{2}\overline{\theta }\theta F(x^\mu ),$$
(18)
where $`\varphi `$ and $`F`$ are real bosonic (even) scalar fields and $`\psi _\alpha `$ is a Majorana spinor field. The SUSY sine-Gordon equation is:
$$\overline{D}D\mathrm{\Phi }=2i\mathrm{sin}\mathrm{\Phi },$$
(19)
where $`D_\alpha =_{\theta ^\alpha }+i(\gamma ^\mu \theta )_\alpha _\mu `$ and on the components it has the form:
$`(\gamma ^\mu _\mu +\mathrm{cos}\varphi )\psi =0`$ (20)
$`\varphi _{xx}\varphi _{tt}={\displaystyle \frac{1}{2}}(\mathrm{sin}(2\varphi )i\overline{\psi }\psi \mathrm{sin}\varphi ).`$ (21)
## IV Super-Hirota operators
In order to apply the bilinear formalism on these equations one has to define a SUSY bilinear operator. We are going to consider the following general $`๐ฉ=1`$ SUSY bilinear expression
$$S_N(f,g)=\underset{i=0}{\overset{N}{}}c_i(D^{Ni}f)(D^ig),$$
(22)
for any N, where $`D`$ is the covariant derivative and $`f`$, $`g`$ are Grassmann valued functions (odd or even). We shall prove the following
Theorem: The general $`๐ฉ=1`$ SUSY bilinear expression (22) is super-gauge invariant i.e. for $`\mathrm{\Theta }=kx+\omega t+\theta \widehat{\zeta }+`$โฆlinears ($`\zeta `$ is a Grassmann parameter)
$$S_N(e^\mathrm{\Theta }f,e^\mathrm{\Theta }g)=e^{2\mathrm{\Theta }}S_N(f,g),$$
if and only if
$$c_i=c_0(1)^{i|f|+\frac{i(i+1)}{2}}\left[\begin{array}{c}N\\ i\end{array}\right],$$
where the super-binomial coefficients are defined by:
$$\left[\begin{array}{c}N\\ i\end{array}\right]=\{\begin{array}{cc}\left(\begin{array}{c}\text{[N/2]}\\ \text{[i/2]}\end{array}\right)\hfill & \text{if }(N,i)(0,1)mod2\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array}$$
$`|f|`$ is the Grassmann parity of the function $`f`$ defined by:
$$|f|=\{\begin{array}{cc}1\hfill & \text{if }f\text{ is odd (fermionic)}\hfill \\ 0\hfill & \text{if }f\text{ is even (bosonic)}\hfill \end{array}$$
and $`[k]`$ is the integer part of the real number $`k`$ ($`[k]k<[k]+1`$)
Proof: First we are going to consider $`N`$ even and we shall take it on the form $`N=2P`$. In this case we have:
$$S_N(f,g)=\underset{i=1}{\overset{N}{}}c_i(D^{Ni}f)(D^ig)=\underset{i=0}{\overset{P}{}}c_{2i}(^{Pi}f)(^ig)+\underset{j=0}{\overset{P1}{}}c_{2j+1}(^{Pj1}Df)(^jDg)$$
Imposing the super-gauge invariance and expanding the covariant derivatives we obtain:
$$\underset{n0}{}\underset{m0}{}\left(\underset{i=0}{\overset{P}{}}c_{2i}\left(\begin{array}{c}i\\ n\end{array}\right)\left(\begin{array}{c}Pi\\ m\end{array}\right)k^{Pnm}\right)(^mf)(^ng)+$$
$$+\underset{n^{}0}{}\underset{m^{}0}{}\mathrm{\Lambda }\left(\underset{j=0}{\overset{P1}{}}c_{2j+1}\left(\begin{array}{c}j\\ n^{}\end{array}\right)\left(\begin{array}{c}Pj1\\ m^{}\end{array}\right)k^{Pn^{}m^{}1}\right)(^m^{}f)(^n^{}Dg)+$$
$$+\underset{n^{}0}{}\underset{m^{}0}{}\mathrm{\Lambda }(1)^{|f|+1}\left(\underset{j=0}{\overset{P1}{}}c_{2j+1}\left(\begin{array}{c}j\\ n^{}\end{array}\right)\left(\begin{array}{c}Pj1\\ m^{}\end{array}\right)k^{Pn^{}m^{}1}\right)(^m^{}Df)(^n^{}g)+$$
$$+\underset{n0}{}\underset{m0}{}\left(\underset{j=0}{\overset{P1}{}}c_{2j+1}\left(\begin{array}{c}j\\ n\end{array}\right)\left(\begin{array}{c}Pj1\\ m\end{array}\right)k^{Pnm1}\right)(^mDf)(^nDg)=$$
$$=\underset{i=0}{\overset{P}{}}c_{2i}(^{Pi}f)(^ig)+\underset{j=0}{\overset{P1}{}}c_{2j1}(^{Pj1}Df)(^jDg)$$
where $`\mathrm{\Lambda }=\widehat{\zeta }+\theta k`$. From this, we must have for every $`m`$, $`n`$ subjected to $`0niPm`$ and $`jPm^{}`$.
$$\underset{i=0}{\overset{P}{}}c_{2i}\left(\begin{array}{c}i\\ n\end{array}\right)\left(\begin{array}{c}Pi\\ m\end{array}\right)k^{Pnm}=c_{2n}\delta _{Pnm}$$
(23)
Also due to the fact that the supergauge invariance has to be obeyed for every $`f`$ and $`g`$ we must have $`c_{2j+1}=0`$ The discrete equation (23) was solved in . Its general solution is given by:
$`c_{2i}`$ $`=`$ $`c_0(1)^i\left(\begin{array}{c}P\\ i\end{array}\right)`$ (26)
$`c_{2j+1}`$ $`=`$ $`0`$ (27)
In the case of $`N=2P+1`$ we proceed in a similar manner and we obtain the following system:
$$\underset{i=0}{\overset{P}{}}c_{2i}\left(\begin{array}{c}i\\ n\end{array}\right)\left(\begin{array}{c}Pi\\ m\end{array}\right)k^{Pnm}=c_{2n}\delta _{Pnm}$$
(28)
$$\underset{j=0}{\overset{P}{}}c_{2j+1}\left(\begin{array}{c}j\\ n\end{array}\right)\left(\begin{array}{c}Pj1\\ m\end{array}\right)k^{Pnm1}=c_{2n+1}\delta _{Pnm1}$$
(29)
$$(1)^{|f|}c_{2i}+c_{2i+1}=0$$
(30)
This system has the following solution:
$`c_{2i}`$ $`=`$ $`c_0(1)^i\left(\begin{array}{c}P\\ i\end{array}\right)`$ (33)
$`c_{2i+1}`$ $`=`$ $`c_0(1)^{i+1+|f|}\left(\begin{array}{c}P\\ i\end{array}\right)`$ (36)
The relations (26), (33) can be written in a compact form as
$$c_i=c_0(1)^{i|f|+\frac{i(i+1)}{2}}\left[\begin{array}{c}N\\ i\end{array}\right].$$
and the theorem is proved. We mention that the super-bilinear operator proposed by McArthur and Yung is a particular case of the above super-Hirota operator.
We shall note the bilinear operator as
$$S_N(f,g):=๐_x^Nfg$$
In the Appendix 1 we list several simple properties of this super-Hirota operator.
## V Bilinear SUSY KdV-type equations
SUSY KdV of Manin-Radul-Mathieu
In order to use the super-bilinear operators defined above we shall consider the following nonlinear substitution for the superfield:
$$\mathrm{\Phi }(t,x,\theta )=2D^3\mathrm{log}\tau (t,x,\theta )$$
(37)
where $`\tau `$ is an even superfield. Introducing in SUSY KdV (11) and integrating with respect to $`x`$ we obtain the following
$$2D_t\mathrm{log}\tau +3\{(2D^3\mathrm{log}\tau )(2_x^2\mathrm{log}\tau )\}+2D^7\mathrm{log}\tau =0$$
Using the properties (62), (63), (64) we find
$$\frac{๐_x๐_t\tau \tau }{\tau ^2}+3\left(\frac{๐_x^3\tau \tau }{\tau ^2}\frac{๐_x^2\tau \tau }{\tau ^2}\right)+2_x^3(\frac{D\tau }{\tau })=0$$
(38)
Now we are using the following property of the classical Hirota operator
$$_x^3(\frac{a}{b})=\frac{๐_x^3ab}{b^2}3\frac{๐_xab}{b^2}\frac{๐_x^2bb}{b^2}$$
In our case $`a=D\tau `$ and $`b=\tau `$. Accordingly
$$2_x^3\left(\frac{D\tau }{\tau }\right)=\frac{๐_x^7\tau \tau }{\tau ^2}3\left(\frac{๐_x^3\tau \tau }{\tau ^2}\frac{๐_x^2\tau \tau }{\tau ^2}\right)$$
Plugging into (38) we find the following super-bilinear form
$$(๐_x๐_t+๐_x^7)\tau \tau =0,$$
(39)
which is equivalent with the form found by McArthur and Yung
$$๐_x(๐_t+๐_x^3)\tau \tau =0.$$
(40)
In order to find the super-soliton solutions we are going to use the classical perturbative method namely the series
$$\tau =1+ฯตf^{(1)}+ฯต^2f^{(2)}+ฯต^3f^{(3)}+\mathrm{}$$
(41)
where $`f^{(i)}`$ are even functions. Equating the power of $`ฯต`$ we find:
* for $`ฯต`$,
$$D(_t+_x^3)f^{(1)}=0$$
(42)
* for $`ฯต^2`$,
$$2D(_t+_x^3)f^{(2)}=๐_x(๐_t+๐_x^3)f^{(1)}f^{(1)}.$$
(43)
* for $`ฯต^3`$,
$$D(_t+_x^3)f^{(3)}=๐_x(๐_t+๐_x^3)f^{(1)}f^{(2)}.$$
(44)
* for $`ฯต^4`$,
$$2D(_t+_x^3)f^{(4)}=2๐_x(๐_t+๐_x^3)f^{(1)}f^{(3)}๐_x(๐_t+๐_x^3)f^{(2)}f^{(2)}$$
(45)
and so on.
Now if we take $`f^{(1)}=e^{kxk^3t+\theta \widehat{\zeta }+\eta ^{(0)}}`$ the equation (42) is satisfied, $`f^{(2)}=0`$, $`f^{(3)}=0\mathrm{}.`$ and the series (41) truncates. So, the 1 supersoliton solution is given by
$$\tau ^{(1)}=1+e^{kxk^3t+\theta \widehat{\zeta }+\eta ^{(0)}}$$
(46)
for every $`k`$ and $`\widehat{\zeta }`$.
Introducing in the super-bilinear equation the 1 soliton solution $`F=1+\mathrm{exp}(kx+\omega t+\widehat{\zeta }\theta )`$ one finds the dispersion supermanifold equation:
$$P(k,\omega ,\widehat{\zeta })(\widehat{\zeta }+\theta k)(\omega +k^3)=0$$
which imposes $`\omega =k^3`$ for every $`\widehat{\zeta }`$.
In order to find 2 super-soliton solution we take $`f^{(1)}=e^{\eta _1}+e^{\eta _2}`$ where $`\eta _i=k_ixk_i^3t+\theta \widehat{\zeta }_i`$ The equation (43) becomes
$$2D(_t+_x^3)f^{(2)}=6k_1k_2(k_1k_2)[(\widehat{\zeta }_1\widehat{\zeta }_2)+\theta (k_1k_2)]e^{\eta _1+\eta _2}$$
(47)
Taking into account that $`\tau `$ is a Grassmann even function, $`f^{(2)}`$ must be also even. Accordingly, the general solution of (47) has the form
$$f^{(2)}=\left[m_{12}(k_1,k_2,\widehat{\zeta }_1,\widehat{\zeta }_2)+\theta \widehat{n}_{12}(k_1,k_2,\widehat{\zeta }_1,\widehat{\zeta }_2)\right]e^{\eta _1+\eta _2},$$
$`m_{12}`$ and $`\widehat{n}_{12}`$ being Grassmann valued even and odd functions. Due to the fact that we have two Grassmann parameters ($`\widehat{\zeta }_1`$ and $`\widehat{\zeta }_2`$) the most general expressions for $`m_{12}`$ and $`\widehat{n}_{12}`$ are given by,
$$m_{12}=(k_1k_2)^2/(k_1+k_2)^2+\gamma (k_1,k_2)\widehat{\zeta }_1\widehat{\zeta }_2$$
$$\widehat{n}_{12}=a(k_1,k_2)\widehat{\zeta }_1+b(k_1,k_2)\widehat{\zeta }_2$$
The above forms for $`m_{12}`$ and $`\widehat{n}_{12}`$ could also be obtained by expanding in power series of $`\widehat{\zeta }_1`$ and $`\widehat{\zeta }_2`$ and taking into account that in the bosonic limit ($`\widehat{\zeta }_i0`$), $`m_{12}(k_1k_2)^2/(k_1+k_2)^2`$ (ordinary KdV interaction term) and $`\widehat{n}_{12}0`$
Introducing in the equation (47) we find
$$f^{(2)}=\left[\left(\frac{k_1k_2}{k_1+k_2}\right)^2+2\frac{k_1k_2}{(k_1+k_2)^2}\widehat{\zeta }_1\widehat{\zeta }_2+2\theta \frac{(k_1k_2)(k_1\widehat{\zeta }_2k_2\widehat{\zeta }_1)}{(k_1+k_2)^2}\right]e^{\eta _1+\eta _2}$$
Introducing the above forms for $`f^{(1)}`$ and $`f^{(2)}`$ in (44) one obtains $`f^{(3)}=0`$ and then $`f^{(4)}=0`$, $`f^{(5)}=0\mathrm{}`$ i.e. the series truncates. So the 2 super-soliton solution is given by;
$$\tau ^{(2)}=1+e^{\eta _1}+e^{\eta _2}+A_{12}e^{\eta _1+\eta _2}$$
(48)
where
$$A_{12}=\left(\frac{k_1k_2}{k_1+k_2}\right)^2+2\frac{k_1k_2}{(k_1+k_2)^2}\widehat{\zeta }_1\widehat{\zeta }_2+2\theta \frac{(k_1k_2)(k_1\widehat{\zeta }_2k_2\widehat{\zeta }_1)}{(k_1+k_2)^2}$$
(49)
One can easily see that the classical general procedure for finding the interaction term given by the equation (7)
$$A_{12}P(k_1+k_2)+P(k_1k_2)=0$$
does not work in the SUSY case.
In order to find 3 supersoliton solution we consider
$$f^{(1)}=e^{\eta _1}+e^{\eta _2}+e^{\eta _3}$$
and the equation for $`f^{(3)}`$ (44)becomes:
$$D(_t+_x^3)f^{(3)}=(๐_x๐_t+๐_x^7)(e^{\eta _1}A_{23}e^{\eta _2+\eta _3}+e^{\eta _2}A_{13}e^{\eta _1+\eta _3}+e^{\eta _3}A_{12}e^{\eta _1+\eta _2})$$
(50)
The solution $`f^{(3)}`$ of this equation (which is very complicated) does not cancel the right hand side of the equation (45). So, $`f^{(4)}`$ is not zero and the series (41) cannot be truncated. Accordingly the SUSY KdV equation does not have 3 super-soliton solution in the standard form for arbitrary $`k_i`$โs and $`\widehat{\zeta }_i`$โs. This seems a quite strange paradox, because SUSY KdV is integrable in the sense of Lax.
Anyway, imposing the constraint $`k_i\widehat{\zeta }_j=k_j\widehat{\zeta }_i`$ for every $`i`$ and $`j`$ it is easy to prove (see Appendix 2) that SUSY KdV possesess the following N-soliton solution
$$\tau ^{(N)}=\underset{\mu =0,1}{}\mathrm{exp}(\underset{i=1}{\overset{N}{}}\mu _i\eta _i+\underset{i<j}{}A_{ij}\mu _i\mu _j),$$
(51)
where
$$\eta _i=k_ixk_i^3t+\theta \widehat{\zeta }_i+\eta _i^{(0)}$$
$$\mathrm{exp}A_{ij}=\left(\frac{k_ik_j}{k_i+k_j}\right)^2$$
$$k_i\widehat{\zeta }_j=k_j\widehat{\zeta }_i$$
Solutions with constraints on parameters have been found also by Liu and Manas , using SUSY Darboux transformation.
SUSY KdV-B
This situation is completely different in the SUSY KdV-B case
$$\mathrm{\Phi }_t+D^6\mathrm{\Phi }+6D^2\mathrm{\Phi }D\mathrm{\Phi }=0.$$
(52)
With the same nonlinear substitution
$$\mathrm{\Phi }=2D^3\mathrm{log}F$$
we obtain the ordinary form
$$(๐_t๐_x+๐_x^4)FF=0$$
which has N-super-soliton solution (5) the fermionic contribution being only an additive phase i.e. $`\eta _i=k_ixk_i^3t+\widehat{\zeta }_i\theta .`$
SUSY Sawada-Kotera-Ramani
For $`๐ฉ=1`$ SUSY Sawada-Kotera-Ramani (15) using the nonlinear substitution,
$$\mathrm{\Phi }=2D^3\mathrm{log}\tau (t,x,\theta )$$
we shall find the following super-bilinear form:
$$(๐_x๐_t+๐_x^{11})\tau \tau =0$$
(53)
In a similar way we find the 2 super-soliton solution
$$\tau ^{(2)}=1+e^{\eta _1}+e^{\eta _2}+[\left(\frac{k_1k_2}{k_1+k_2}\right)^2\frac{k_1^2k_1k_2+k_2^2}{k_1^2+k_1k_2+k_2^2}+2\frac{k_1k_2}{(k_1+k_2)^2}\frac{k_1^2k_1k_2+k_2^2}{k_1^2+k_1k_2+k_2^2}\widehat{\zeta }_1\widehat{\zeta }_2]\times $$
$$\times (1+2\theta \frac{k_2\widehat{\zeta }_1k_1\widehat{\zeta }_2}{k_1k_2})e^{\eta _1+\eta _2}$$
Also with the same constraint $`k_i\widehat{\zeta }_j=k_j\widehat{\zeta }_i`$ we find
$$\tau ^{(N)}=\underset{\mu =0,1}{}\mathrm{exp}(\underset{i=1}{\overset{N}{}}\mu _i\eta _i+\underset{i<j}{}A_{ij}\mu _i\mu _j),$$
(54)
where
$$\eta _i=k_ixk_i^5t+\theta \widehat{\zeta }_i+\eta _i^{(0)}$$
$$\mathrm{exp}A_{ij}=\left(\frac{k_ik_j}{k_i+k_j}\right)^2\frac{k_i^2k_ik_j+k_j^2}{k_i^2+k_ik_j+k_j^2}$$
SUSY Hirota-Satsuma
For $`๐ฉ=1`$ SUSY Hirota-Satsuma equation using the nonlinear substitution:
$$\mathrm{\Phi }=2D\mathrm{log}\tau (t,x,\theta )$$
one obtains the super-bilinear form:
$$(๐_x^5๐_t๐_x^3๐_x๐_t)\tau \tau =0$$
(55)
The 2 supersoliton solution is given by:
$$\tau ^{(2)}=1+e^{\eta _1}+e^{\eta _2}+\left[\left(\frac{k_1k_2}{k_1+k_2}\right)^2M_{12}+2\frac{k_1k_2}{(k_1+k_2)^2}M_{12}\widehat{\zeta }_1\widehat{\zeta }_2\right](1+2\theta \frac{k_2\widehat{\zeta }_1k_1\widehat{\zeta }_2}{k_1k_2})e^{\eta _1+\eta _2}$$
where
$$M_{12}=\frac{(k_1k_2)^2+k_1k_2[(k_1k_2)^2(k_1^21)(k_2^21)]}{(k_1k_2)^2k_1k_2[(k_1k_2)^2(k_1^21)(k_2^21)]}$$
and $`\eta _i=k_ixk_it/(k_i^21)+\widehat{\zeta }_i\theta .`$ With the constraints $`k_i\widehat{\zeta }_j=k_j\widehat{\zeta }_i`$ the equation admits also N soliton solution with
$$A_{ij}=\left(\frac{k_1k_2}{k_1+k_2}\right)^2M_{12}.$$
SUSY Burgers
An interesting case is the SUSY extension of the Burgers equation, namely
$$\mathrm{\Phi }_t+\mathrm{\Phi }D\mathrm{\Phi }_x+\mathrm{\Phi }_{xx}=0$$
It is welknown that the classical Burgers equation can be linearized via Cole-Hopf transform. If we try the nonlinear substitution (which is the natural supersymmetrization of the Cole-Hopf transform):
$$\mathrm{\Phi }=2D\mathrm{log}\tau (t,x,\theta )$$
we find:
$$๐_x๐_t\tau \tau +2๐_x^2D\tau \tau =0.$$
This form is not super-gauge invariant. Accordingly we are forced to use two functions for substitution, namely
$$\mathrm{\Phi }=\frac{\widehat{G}(t,x,\theta )}{F(t,x,\theta )}$$
where $`\widehat{G}`$ is an odd Grassmann function and $`F`$ is an even one. Using the relation (65) from Appendix 1 we obtain the following gauge-invariant super-bilinear form
$$(๐_t+๐_x^2)\widehat{G}F=0$$
$$๐_x^2FF=๐_x^3\widehat{G}F$$
This system admits the following 1 super-shock solution:
$$\widehat{G}=2(\widehat{\zeta }+\theta k)e^{(kxk^2t+\widehat{\zeta }\theta )},F=1+e^{(kxk^2t+\widehat{\zeta }\theta )}$$
We can ask ourselves if it is possible to obtain super-bilinear forms for SUSY equations of the nonlinear Klein-Gordon type. In fact the SUSY sine-Gordon equation(19) can be written in the following form:
$$[D_T,D_X]\mathrm{\Phi }(T,X,\theta ,\theta _t)=2i\mathrm{sin}\mathrm{\Phi }(T,X,\theta ,\theta _t)$$
(56)
where we have introduced the light-cone variables $`X:=i(tx)/2`$, $`T:=i(t+x)/2`$, and $`\theta :=\theta _1`$, $`\theta _2:=\theta _t.`$ Covariant derivatives are $`D_X:=_\theta +\theta _X`$, $`D_T:=_{\theta _t}+\theta _t_T`$ and the square braket means the commutator. Using the nonlinear substitution ($`G`$ and $`F`$ are even functions)
$$\mathrm{\Phi }=2i\mathrm{log}\left(\frac{G}{F}\right),$$
we find the following quadrilinear expression
$$2i\{F^2(G[D_T,D_X]G[D_TG,D_XG])G^2(F[D_T,D_X]F[D_TF,D_XF])\}=F^4G^4$$
It is easy to see that the bilinear operator
$$๐_{XT}\tau \tau :=\tau [D_T,D_X]\tau [D_T\tau ,D_X\tau ]$$
is super-gauge invariant with respect to the super-gauge
$$e^\mathrm{\Theta }:=e^{(kx+\omega t+\theta \widehat{\zeta }+\theta _t\widehat{\mathrm{\Omega }}+liniars)}.$$
Accordingly we can choose the following super-bilinear form, formally the same with standard sine-Gordon equation,
$`๐_{XT}GG`$ $`=`$ $`{\displaystyle \frac{1}{2i}}(F^2G^2)`$ (57)
$`๐_{XT}FF`$ $`=`$ $`{\displaystyle \frac{1}{2i}}(G^2F^2)`$ (58)
but, it is not clear how to compute the super-kink solutions.
ยฟFrom these examples it seems that gauge-invariance is a useful concept for bilinear formalism in the supersymmetric case, though there is no deep reason for that. As a consequence we was able to bilinearize two supersymmetric equations of KdV type and the SUSY sine-Gordon. The case of SUSY versions for mKdV, NLS, KP etc. requires further investigation because it seems that only certain supersymmetric extensions are super-bilinearizable. Although we do not know if the SUSY extension of Sawada-Kotera proposed above are integrable in the sense of Lax, it admits super-bilinear form and only 2 super-soliton solution for arbitrary choice of solitary waves. The strange fact is that SUSY KDV equation of Mathieu which is known to be Lax integrable does not admit $`Ngeq3`$ supersoliton solution in the canonical form. Probably a singularity analysis implemented on the super-bilinear form will reveal the connection between Hirota-integrability and Lax-integrability.
## VI Appendix 1
In this section we are going to list several properties of the super-Hirota bilinear operator which are useful in deriving bilinear forms.
$$๐_x^{2N}fg=๐_x^Nfg$$
(59)
$$๐_x^{2N+1}e^{\eta _1}e^{\eta _2}=[\widehat{\zeta }_1\widehat{\zeta }_2+\theta (k_1k_2)](k_1k_2)^Ne^{\eta _1+\eta _2}$$
(60)
$$๐_x^{2N+1}1e^{\eta _1}=(1)^{N+1}(\widehat{\zeta }+\theta k)k^Ne^\eta =(1)^{N+1}๐_x^{2N+1}e^\eta 1$$
(61)
where $`\eta _i=k_ix+\theta \widehat{\zeta }_i`$ and $`\widehat{\zeta }_i`$ are odd Grassmann numbers.
$$2D\mathrm{log}\tau =\frac{D\tau }{\tau }$$
(62)
$$2D^3\mathrm{log}\tau =\frac{๐_x^3\tau \tau }{\tau ^2}$$
(63)
$$2D_t\mathrm{log}\tau =\frac{๐_x๐_t\tau \tau }{\tau ^2}$$
(64)
where $`\tau `$ is an even Grassmann function Moreover if $`G`$ and $`F`$ are Grassmann functions (with F even) then
$$D^3(\frac{G}{F})=\frac{๐_x^3GF}{F^2}(1)^{|G|}\frac{G}{F}\frac{๐_x^3FF}{F^2}$$
(65)
## VII Appendix 2
In this section we are going to sketch the proof of the formula for N supersoliton solution for bilinear SUSY KdV equation. We are rely on the proof of Hirota for ordinary N-soliton solution in the case of KdV , . Thus, introducing the expression of the N supersoliton solution
$$\tau ^{(N)}=\underset{\mu =0,1}{}\mathrm{exp}(\underset{i=1}{\overset{N}{}}\mu _i\eta _i+\underset{i<j}{}A_{ij}\mu _i\mu _j),$$
$$\eta _i=k_ixk_i^3t+\theta \widehat{\zeta }_i+\eta _i^{(0)}$$
$$\mathrm{exp}A_{ij}=\left(\frac{k_ik_j}{k_i+k_j}\right)^2$$
$$k_i\widehat{\zeta }_j=k_j\widehat{\zeta }_i$$
into the super bilinear form
$$(๐_x๐_t+๐_x^7)\tau \tau =0,$$
and taking into account the properties (59) and (60) we find
$$\underset{\mu =0,1}{}\underset{\nu =0,1}{}\{(\underset{i=1}{\overset{N}{}}(\mu _i\nu _i)\mathrm{\Lambda }_i)(()\underset{i=1}{\overset{N}{}}(\mu _i\nu _i)k_i^3)+(\underset{i=1}{\overset{N}{}}(\mu _i\nu _i)\mathrm{\Lambda }_i)(\underset{i=1}{\overset{N}{}}(\mu _i\nu _i)k_i)^3\}\times $$
$$\times \mathrm{exp}(\underset{i=1}{\overset{N}{}}(\mu _i+\nu _i)\eta _i+\underset{i<j}{}(\mu _i\mu _j+\nu _i\nu _j)A_{ij})=0$$
where $`\mathrm{\Lambda }=\widehat{\zeta }_i+\theta k_i`$. Since $`\mu _i,\nu _i=0,1`$ it is clear that we only have exponential terms of the form
$$\mathrm{exp}(\underset{i=1}{\overset{n}{}}\eta _i+\underset{i=n+1}{\overset{m}{}}2\eta _i),0nN,nmN$$
Next we show that the coefficient of this general exponential term is zero, the coefficient being given by, :
$$\mathrm{\Delta }=\underset{\sigma =0,1}{}\{(\underset{i=1}{\overset{N}{}}\sigma _i\mathrm{\Lambda }_i)(\underset{i=1}{\overset{N}{}}\sigma _ik_i^3)+(\underset{i=1}{\overset{N}{}}\sigma _i\mathrm{\Lambda }_i)(\underset{i=1}{\overset{N}{}}\sigma _ik_i)^3\}\underset{i<j}{\overset{n}{}}(\sigma _ik_i\sigma _jk_j)^2$$
(66)
where $`\sigma _i=\mu _i\nu _i`$. We do not go into details because the procedure of deriving the above coefficient goes absolutely in the same way as in the case of ordinary KdV equation. This is due to the fact that the interaction term $`A_{ij}`$ is the same for KdV and SUSY KdV.
Because $`\mathrm{\Lambda }=\widehat{\zeta }_i+\theta k_i`$ the coefficient $`\mathrm{\Delta }`$ becomes on the components
$$\mathrm{\Delta }\mathrm{\Delta }_0+\mathrm{\Delta }_1=\underset{\sigma =0,1}{}\{(\underset{i=1}{\overset{N}{}}\sigma _i\widehat{\zeta }_i)(\underset{i=1}{\overset{N}{}}\sigma _ik_i^3)+(\underset{i=1}{\overset{N}{}}\sigma _i\widehat{\zeta }_i)(\underset{i=1}{\overset{N}{}}\sigma _ik_i)^3\}\underset{i<j}{\overset{n}{}}(\sigma _ik_i\sigma _jk_j)^2+$$
(67)
$$+\theta \underset{\sigma =0,1}{}\{(\underset{i=1}{\overset{N}{}}\sigma _ik_i)(\underset{i=1}{\overset{N}{}}\sigma _ik_i^3)+(\underset{i=1}{\overset{N}{}}\sigma _ik_i)(\underset{i=1}{\overset{N}{}}\sigma _ik_i)^3\}\underset{i<j}{\overset{n}{}}(\sigma _ik_i\sigma _jk_j)^2$$
Hirota proved that the second term is zero , . The first term is also zero because, using the property $`\widehat{\zeta }_ik_j=\widehat{\zeta }_jk_i`$, it can be written as
$$\mathrm{\Delta }_0=\frac{\widehat{\zeta }_m}{k_m}\mathrm{\Delta }_1=0$$
|
warning/0006/nucl-th0006030.html
|
ar5iv
|
text
|
# Dense Nuclear Matter: Landau Fermi-Liquid Theory and Chiral Lagrangian with Scaling
## 1 Introduction
Although QCD which deals with quarks and gluons is believed to be the fundamental theory for strong interactions, it is generally accepted that the appropriate theory at very low energy is the effective quantum field theory which incorporates the observed degrees of freedom in low-energy nuclear process, i.e., pions, nucleons and other low-mass hadrons . The effective Lagrangian in matter-free or dilute space is governed by QCD symmetries with its parameters to be determined from experiments in free space. The energy scale of the experiments in which one is interested determines which hadrons play an important role in the theory. For example, it is shown that one can integrate out even the pions for the two nucleon systems at very low energy. According to the results of , the deuteron and low-energy nucleon-nucleon scattering properties can be described very accurately by an effective theory given in terms of the nuclonic degrees of freedom only with a cutoff around the natural scale of the theory which for low energy is the pion mass.
Since there is currently growing interplay between the physics of hadrons and the physics of compact objects in astrophysics through the properties of hot and dense environments, we want to extend such successful strategy of effective field theories to a dense medium. First of all, our main goal is to understand the properties of dense hadronic medium which can be tested in various heavy ion collisions. Recent dilepton experiments (CERES and HELIOS-3) gave us very important information on the properties of hadrons in dense medium, which we will detail later. Understanding the properties of dense hadronic medium through heavy ion collision experiments is essential for understanding the properties of neutron stars which are believed to be formed in the center of supernovae at the time of explosions. Especially the determination of the maximum neutron star mass is one of the most important issues in astrophysics in explaining controversies between the observations and theoretical estimations, and it will give some hints for the detectibility of the gravitational wave detectors. A dense matter makes the probed energy scale larger than that in free space. Therefore we have to introduce more massive degrees of freedom which are usually vector meson and/or higher order operators in the nucleon fields. We must also consider a new energy scale, Fermi energy of nucleons in bound system.
The standard strategy to attack the dense hadronic matter is to obtain the ground state of matter and compute the excitations around it, based on an effective Lagrangian whose parameters are obtained in free space. Although such an approach can give satisfactory results with a sufficient number of parameters, it is not obvious whether we can extend the results for the different density. When the extension does not work, we usually face with very complicated loop diagrams which may lead to the impasse.
Another strategy is to start from the in-medium Lagrangian which is built on the reasonable assumptions, instead of deriving the hadronic matter properties from the matter-free effective Lagrangian defined in free space. In this approach one regards its mean field solution as a solution of the Wilsonian effective action in which the high energy modes are integrated out into the coefficient. We can compare it with the well-known Landau Fermi-liquid theory, which works at low energy excitation in strongly correlated Fermi system. Landau Fermi-liquid theory is described by quasiparticles, which are the low energy excitations in Fermi liquid, and their interactions under the assumption of one-to-one correspondence between the quasiparticle in the liquid and the particle in a non-interacting gas. It is a fixed point theory under Wilsonian renormalization to the Fermi surface with $`\mathrm{\Lambda }/k_F0`$ if Bardeen-Cooper-Schrieffer(BCS) instability does not exist, where $`\mathrm{\Lambda }`$ is the cutoff of the theory relative to the Fermi surface and $`k_F`$ is Fermi momentum. Since the result after the repeated renormalization does not depend on $`k_F`$, the argument for Fermi liquid holds as long as there is no phase transition. A famous example of such a Lagrangian is Walecka model. Its extension and justification were studied recently by Furnstahl et al. Their Lagrangian is constrainted by QCD symmetry and by Georgiโs naturalness condition and the coupling constants in it are tuned in order to describe nuclear ground state. Bulk properties of nuclei are described by it very successfully. But it is somewhat unclear to know how to approach the fluctuation on the ground state.
In this review we will approach the nuclear matter in a different way. We wish to apply the strategy of the effective field theory to the in-medium theory. We assume that the in-medium effective Lagrangian has the same structure as in free space according to the symmetry constraint of the fundamental theory QCD but that its parameters are modified in medium. It means that the effect of embedding a hadron in matter appears mainly in the change of the โvacuum,โ i.e., quark and gluon condensate in QCD variables and the parameters in an effective theory. In this strategy the density-dependent parameters include many-body correlations.
Brown-Rho(BR) scaling is one specific way to define such in-medium parameters. Brown-Rho scaling is the scaling of the dynamically generated masses of hadrons which consist of chiral quarks, i.e., $`u`$ and $`d`$ quarks. Brown and Rho phrased the scaling with the large $`N_c`$ Lagrangian, i.e., Skyrmion, under the assumption that the chiral symmetry and the scale symmetry of QCD are relevant. Their Lagrangian is implemented with scale anomaly of QCD, too. The masses and pion decay constant of this QCD effective theory scale universally:
$`\mathrm{\Phi }(\rho ){\displaystyle \frac{f_\pi ^{}(\rho )}{f_\pi }}{\displaystyle \frac{m_v^{}(\rho )}{m_v}}{\displaystyle \frac{m_\sigma ^{}(\rho )}{m_\sigma }}{\displaystyle \frac{M^{}(\rho )}{M}}.`$ (1)
The star represents in-medium quantities here. $`v`$ is vector meson degree of freedom and $`s`$ an isoscalar scalar meson which has a mass $``$ 500 MeV in nuclear matter. $`M`$ represents a free nucleon mass and $`M^{}`$ a scaled nucleon mass which is somewhat different from the Landau effective mass discussed in Section 6.1.
It is known that BR scaling describes that the light-quark vector meson property in the extreme condition very successfully. One can make the extreme condition through relativistic heavy ion collisions. Specially the dileptons provide a good probe of the earlier dense and hot stage of relativistic heavy ion collision since the interaction of leptons is not subject to the strong interactions of the final state. CERES(Cherenkov Ring Electron Spectrometer) collaboration observed in heavy ion collision (S$`+`$Au) that the dilepton production with invariant masses from 250 MeV to about 500 MeV is enhanced much more than the predicted from the superposition of $`pp`$ collision . And HELIOS-3(CERN Super Proton Synchrotron detector) also observed the dilepton enhancement in S$`+`$W. It is shown by Li, Ko, and Brown that a chiral Lagrangian with BR-scaled meson masses describes most economically and beautifully the enhancement, which is found to come from the dropping vector meson masses in dense matter. Furthermore the excitation into the kaonic direction above the given ground state seems to describe the properties of kaons in medium with scaled parameters .
Since BR scaling gives the universal scaling mass relation among hadrons, it must work for the baryon properties in medium. In recent works it has been discussed how the BR scaling, which is applied to meson properties successfully, can be applied to the baryon property in dense medium and how it can be extrapolated to a hadronic matter under extreme conditions from the known normal nuclear matter. The construction of such bridges will be necessary to understand the various phenomena in relativistic heavy ion collisions and in compact stars in the universe. For those purposes we develop arguments for mapping the effective chiral Lagrangian whose parameters are governed by BR scaling to Landau Fermi-liquid theory. We will relate the meson mass scaling with baryon mass scaling and also relate matter properties with BR scaling parameter, which implies the vacuum structure characterized by quark condensate, via Landau parameter. Fermi liquid theory may work up to chiral phase transition, though the relevant degrees of freedom are changed to quasiquarks from quasihadrons.
In this review we approach the aim in two ways. The first is to obtain the ground state with BR scaling. How the Fermi surface is obtained in BR scaling framework is not yet understood clearly. So people usually assume that the ground state of hadronic matter is determined by the conventional matter from a standard many-body theory and use density-dependent effective chiral Lagrangians to compute mesonic fluctuations above the ground state. Though various fluctuation phenomena can be described successfully in this way, such a treatment gives no constraints for consistency between the excitations and the ground state. Though BR scaling is applied very successfully to describe meson properties in medium , the way to deal with the matter properties is disconnected from BR scaling. Such a procedure is not satisfactory in going to the higher density region from the normal nuclear matter density. We can take the kaon condensation in neutron star as an example. In dealing with it, we come across a change of the ground state from that of a non-strange matter to a strange matter. The works up to date treated $`KN`$ interaction and the ground state separately. This is not satisfactory. Ground state properties might effect the condensation. For example, the effect of the four-Fermi interactions which play an important role in determining ground state, suppresses a pion condensation . So we must deal with the whole bulk involving the ground state and excitations on top of it on the same footing. We assume that the ground state is given by the same effective Lagrangian that is supposed to include higher order corrections as the mean field of the BR-scaled chiral Lagrangian. We want to make an initial step for dealing with the ground state and the fluctuation on top of it on the same basis. We bridge these two properties by constructing a simple model whose parameters scale in the manner of BR and which describes nuclear matter properties well. It is shown that the model can be mapped to Landau Fermi-liquid theory.
The next step to achieve our aim is to identify the parameters of the BR-scaled effective Lagrangian with the fixed point quantities in Landau Fermi-liquid theory, given a hadronic matter with a Fermi surface. With this identification, certain mean field quantities of heavy meson (e.g. $`\rho `$, $`\omega `$) can be related to BR scaling through the Landau parameters. We show how such arguments work for the electromagnetic currents in nuclear matter. We can link a set of BR-scaled parameters at nuclear matter density with the orbital gyromagnetic ratio in terms of the Landau parameter $`F_1^\omega `$ which comes from integrated-out isoscalar vector degree of freedom in the effective Lagrangian. Then we will try to derive the corresponding formulas for the axial current in a similar way.
This review is organized as follows. In Section 2 Landau Fermi-liquid theory and its interpretation in terms of renormalization group language are summarized briefly. And we show how thermodynamic observables are related to the relativistic Landau parameters. In Section 3 the strategy of an effective field theory and how it is applied to a dense matter are explained. We proceed to discuss how nuclear matter described by an in-medium effective Lagrangian can be identified with Landau Fermi liquid. The model of Furnstahl, Serot and Tang (referred to FTS1) which imposes the symmetries of QCD is examined as an example of the application of a general strategy of an effective chiral theory to a medium. In Section 4 Brown-Rho scaling is derived with QCD-oriented effective Lagrangian. A model where BR scaling governs the parameters of a chiral Lagrangian and determines the background at a given density is constructed in Section 5 in order to describe in weak coupling the same physics as FTS1 which has strong coupling in the form of a large anomalous dimension of a dilaton. It describes well normal nuclear matter properties and has thermodynamic consistency and Fermi-liquid structure needed for the extrapolation to higher density region. In Section 6 vector current and axial charge transition matrix elements for a nucleon above the given Fermi sea in Landau-Migdal theory and in chiral effective Lagrangian are calculated and compared. A summary and comments on some unsolved problems are given in Section 7 and Section 8 respectively. Appendix A shows how sensitive the equation of state is to the many-body correlation parameters for $`\rho >\rho _0`$. Appendix B shows how to compute relativistically the pionic contribution to Landau parameter $`F_1`$ by means of Fierz transformation. And the vector-mesonic contribution to $`F_1`$ and to electromagnetic convection current is calculated in Appendix C relativistically with random phase approximation.
## 2 Landau Fermi-liquid theory
We discuss briefly Landau Fermi-liquid theory in this section, before presenting the relation between the chiral effective theory for nuclear matter and Landau Fermi-liquid theory. A mini-primer on Landau Fermi-liquid theory is given in the first subsection to define the quantities involved. In the second subsection, it is discussed that Landau Fermi-liquid theory is considered as an effective theory and is shown to be a fixed point theory in renormalization group(RG) language. And we will show how the thermodynamic quantities are related to relativistic Landau parameters in Section 2.3.
### 2.1 A mini-primer
Landauโs Fermi-liquid theory is a semi-phenomenological approach to strongly interacting normal Fermi systems at small excitation energies. The elementary excitations of the Fermi-liquid, which correspond to single particle degrees of freedom of the Fermi gas, are called quasiparticles in Landau Fermi-liquid theory. It is assumed that a one-to-one correspondence exists between the low-energy excitations of the Fermi liquid near Fermi surface, i.e., quasiparticles, and those of a non-interacting Fermi gas. A quasiparticle state of the interacting liquid is obtained by turning on the interaction adiabatically at the corresponding state of non-interacting Fermi gas. The quasiparticle properties, e.g. the mass, in general differ from those of free particles due to interaction effects. In addition there is a residual quasiparticle interaction, which is parameterized in terms of the so called Landau parameters.
The adiabatic process described above is possible in the vicinity of the Fermi surface only. Let us see the Fermi liquid at $`T=0`$. Since Pauli exclusion principle makes the states below the Fermi surface filled, quasiparticle with energy $`\epsilon `$ loses energy less than $`\epsilon \epsilon _F`$ when colliding the background particles. It means that the quasiparticles which can interact with the quasiparticle are those with an energy within $`|\epsilon \epsilon _F|`$ of the Fermi surface. And the final state momenta are also restricted by $`\epsilon ^{}<\epsilon `$. Pauli exclusion principle and the corresponding rarity of final states make the quasiparticle life time proportional to $`|\epsilon \epsilon _F|^2`$ at $`T=0`$ case.
Fermi-liquid theory is a prototype effective theory, which works because there is a separation of scales. The theory is applicable to low-energy phenomena, while the parameters of the theory are determined by interactions at higher energies. The separation of scales is due to the Pauli principle and the finite range of the interaction. Pauli principle makes the low energy quasiparticle physics possible near the Fermi surface and the finite range of interaction makes a few quasiparticles around Fermi surface, who appear by the small change of the energy in low energy physics, form a gas. Fermi-liquid theory has proven very useful for describing the properties of e.g. liquid <sup>3</sup>He and provides a theoretical foundation for the nuclear shell model as well as nuclear dynamics of low-energy excitations .
The interaction between two quasiparticles $`๐_1`$ and $`๐_2`$ at the Fermi surface of symmetric nuclear matter can be written in terms of a few spin and isospin invariants
$`f_{๐_1\sigma _1\tau _1,๐_2\sigma _2\tau _2}`$ $`=`$ $`{\displaystyle \frac{1}{N(0)}}[F(\mathrm{cos}\theta _{12})+F^{}(\mathrm{cos}\theta _{12})๐_1๐_2+G(\mathrm{cos}\theta _{12})๐_1๐_2`$ (2)
$`+`$ $`G^{}(\mathrm{cos}\theta _{12})๐_1๐_2๐_1๐_2+{\displaystyle \frac{๐^{\mathrm{\hspace{0.17em}2}}}{k_F^2}}H(\mathrm{cos}\theta _{12})S_{12}(\widehat{๐})`$
$`+`$ $`{\displaystyle \frac{๐^{\mathrm{\hspace{0.17em}2}}}{k_F^2}}H^{}(\mathrm{cos}\theta _{12})S_{12}(\widehat{๐})๐_1๐_2]`$
where $`\theta _{12}`$ is the angle between $`๐_1`$ and $`๐_2`$ and $`N(0)=\frac{\gamma k_F^2}{(2\pi ^2)}\left(\frac{dp}{d\epsilon }\right)_F`$ is the density of states at the Fermi surface. In this review natural units where $`\mathrm{}=1`$ are used. The spin and isospin degeneracy factor $`\gamma `$ is equal to 4 in symmetric nuclear matter. Furthermore, $`๐=๐_1๐_2`$ and
$$S_{12}(\widehat{๐})=3๐_1\widehat{๐}๐_2\widehat{๐}๐_1๐_2,$$
(3)
where $`\widehat{๐}=๐/|๐|`$. The tensor interactions $`H`$ and $`H^{}`$ turn out to be important for the axial charge The functions $`F,F^{},\mathrm{}`$ are expanded in Legendre polynomials,
$$F(\mathrm{cos}\theta _{12})=\underset{\mathrm{}}{}F_{\mathrm{}}P_{\mathrm{}}(\mathrm{cos}\theta _{12}),$$
(4)
with analogous expansion for the spin- and isospin-dependent interactions. The energy of a quasiparticle with momentum $`p=|๐|`$, spin $`\sigma `$ and isospin $`\tau `$ is denoted by $`ฯต_{p,\sigma ,\tau }`$ and the corresponding quasiparticle number distribution by $`n_{p,\sigma ,\tau }`$. From now on the spin and isospin indices $`\sigma `$ and $`\tau `$ will be omitted from the formulas to avoid overcrowding, except where needed to avoid ambiguities.
$`n_p(๐,t)`$ $`=`$ $`n_p^0(\epsilon _p^0)+\delta n_P(๐,t)`$ (5)
$`\epsilon _p(๐,t)`$ $`=`$ $`\epsilon _p^0+{\displaystyle \underset{\sigma ^{},\tau ^{}}{}}{\displaystyle \frac{d^3p^{}}{(2\pi )^3}f_{pp^{}}\delta n_p^{}(๐,t)}`$ (6)
where $`\delta n_p(๐,t)`$ is the long wave length excitations from the ground state $`n_p^0(\epsilon ^0)`$ in the vicinity of Fermi surface. The space and time dependence of the quantities will also be omitted, e.g., $`\epsilon \epsilon (๐,t)`$. The Landau effective mass and velocity of a quasiparticle on the Fermi surface is defined by
$$\frac{d\epsilon _p}{dp}|_{p=k_F}=\frac{k_F}{m_L^{}}v_F^{}.$$
(7)
We must note that the total current is not
$`๐ฑ_{locQP}={\displaystyle \underset{\sigma ,\tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}๐_F^{}\delta n_p}`$ (8)
in Landau Fermi-liquid theory. The quasiparticle distribution $`n_p(๐,t)`$ obeys
$`{\displaystyle \frac{n_p}{t}}+{\displaystyle \frac{d๐}{dt}}\mathbf{}_rn_p+{\displaystyle \frac{d๐}{dt}}\mathbf{}_pn_p=0.`$ (9)
The key assumption of Landau Fermi-liquid kinetic theory is that $`\epsilon _p(๐,t)`$ plays the role of the quasiparticle Hamiltonian;
$`{\displaystyle \frac{d๐}{dt}}`$ $`=`$ $`\mathbf{}_r\epsilon _p`$ (10)
$`{\displaystyle \frac{d๐}{dt}}`$ $`=`$ $`\mathbf{}_p\epsilon _p.`$ (11)
The kinetic equation is
$`{\displaystyle \frac{n_p}{t}}+\mathbf{}_rn_p\mathbf{}_p\epsilon _p\mathbf{}_pn_p\mathbf{}_r\epsilon _p=I[n_p^{}]`$ (12)
where $`I[n_p^{}]`$ is an internal collision integral which represents the sudden change of quasiparticle momenta. We consider the system without external forces. Integrating over $`๐`$, that cancels the effect of the internal collision under the assumption that the quasiparticle energy, momentum and number are locally conserved, (12) becomes (to order $`\delta n_p(๐,t)`$), by integration by part,
$`{\displaystyle \underset{\sigma ,\tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}\frac{\delta n_p}{t}}+\mathbf{}_r(n_p\mathbf{}_p\epsilon _p)`$ $`=`$ $`{\displaystyle \frac{\delta \rho }{t}}+{\displaystyle \underset{\sigma ,\tau }{}}\mathbf{}_r{\displaystyle \frac{d^3p}{(2\pi )^3}\delta n_p\mathbf{}_p\epsilon _p}+\delta \epsilon _p\mathbf{}_pn_p^0`$ (13)
$`=`$ $`0`$
where $`\delta \rho `$ is the total particle (or quasiparticle) density fluctuation. We can easily see that the conserved current is
$`๐ฑ`$ $`=`$ $`{\displaystyle \underset{\sigma ,\tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}n_p\mathbf{}_p\epsilon _p}`$ (14)
$`=`$ $`{\displaystyle \underset{\sigma ,\tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}๐_F^{}\delta n_p^{loc}}`$
with the excitation from the local equilibrium $`n_p^0(\epsilon (๐,t))`$;
$`n_p(๐,t)=n_p^0(\epsilon (๐,t))+\delta n_p^{loc}(๐,t).`$ (15)
Comparing (15) with (5) and (6), we obtain
$`\delta n_p^{loc}=\delta n_p+{\displaystyle \frac{n_p^0}{\epsilon }}{\displaystyle \underset{\sigma ^{},\tau ^{}}{}}{\displaystyle \frac{d^3p^{}}{(2\pi )^3}}f_{pp^{}}\delta n_p^{}.`$ (16)
The modification $`๐ฑ_{locQP}๐ฑ`$ can be interpreted as the effect of the return flow of the surrounding matter to the localized wave packet carrying $`๐ฑ`$ and is called back-flow current.
### 2.2 Renormlization group approach
Landau Fermi-liquid theory was based on Landauโs reasonable intuition. After his work, his theory was derived and justified microscopically . Recent development of Wilsonian RG method in medium provides us a new understanding of Fermi liquid theory. Landau Fermi-liquid theory is a fixed point theory described by marginal coupling, i.e. $``$, $`m_L^{}`$ . In this section we will review the RG arguments for Fermi liquid theory.
To do this we consider a nonrelativistic system of spinless fermions whose Fermi surface is spherical characterized by $`k_F`$ for simplicity. Then non-interacting one particle Hamiltonian near Fermi surface is
$`H={\displaystyle \frac{๐ฒ^2}{2m}}{\displaystyle \frac{k_F^2}{2m}}{\displaystyle \frac{k}{m}}k_Fv_Fk`$ (17)
with $`k=|๐ฒ|k_F`$. The free fermion field action
$`S_0={\displaystyle _\mathrm{\Lambda }}\overline{\psi }(\omega k\mathrm{\Omega })(i\omega vk)\psi (\omega k\mathrm{\Omega })`$ (18)
in momentum space where
$`{\displaystyle _\mathrm{\Lambda }}:={\displaystyle \frac{d\mathrm{\Omega }}{(2\pi )^2}_\mathrm{\Lambda }^\mathrm{\Lambda }\frac{dk}{(2\pi )}_{\mathrm{}}^{\mathrm{}}\frac{d\omega }{(2\pi )}}.`$ (19)
$`\overline{\psi }`$ and $`\psi `$ are Grassmannian eigenvalue with fermion operator $`\widehat{\mathrm{\Psi }}`$; $`\widehat{\mathrm{\Psi }}|\psi =\psi |\psi `$ and $`\overline{\psi }|\widehat{\mathrm{\Psi }}^{}=\overline{\psi }|\overline{\psi }`$. Note that a shell of thickness $`\mathrm{\Lambda }`$ on either side of the Fermi surface is taken for low energy physics as seen in Fig. 1.
Pauli exclusion principle lets only small deviations from the Fermi surface, not from the origin, be important in low energy physics of fermion matter. This defines the starting point of an in-medium renormalization group procedure.
The first step for the renormalization is decimation; to integrate out the high energy mode whose momentum is larger than $`\mathrm{\Lambda }/s`$ and to reduce the cutoff from $`\mathrm{\Lambda }`$ to $`\mathrm{\Lambda }/s`$. For example, the free action (18) becomes
$`S_0^{eff}={\displaystyle \frac{d\mathrm{\Omega }}{(2\pi )^2}_{\mathrm{\Lambda }/s}^{\mathrm{\Lambda }/s}\frac{dk}{(2\pi )}_{\mathrm{}}^{\mathrm{}}\frac{d\omega }{(2\pi )}\overline{\psi }(\omega k\mathrm{\Omega })(i\omega vk)\psi (\omega k\mathrm{\Omega })}.`$ (20)
Next we rescale the momenta in order to compare the old and the new;
$`(\omega ,๐)(s\omega ,s๐).`$ (21)
The last step is to absorb the uninteresting multiplicative constant by
$`\psi s^{3/2}\psi .`$ (22)
The RG transformation consists of these three steps. After such RG transformation, the free action (18) returns to the old. When a coupling is turned on, the coupling is called relevant if it increases after RG transformation. If it decreases, it is called irrelevant and if it remains fixed like the free action, it is called marginal.
Now let us turn on four-Fermi interaction. The appropriate action is
$`S`$ $`=`$ $`{\displaystyle _\mathrm{\Lambda }}\overline{\psi }[i\omega v_F^{}k]\psi +\delta \mu ^{}{\displaystyle _\mathrm{\Lambda }}\overline{\psi }\psi `$ (23)
$`+{\displaystyle \frac{1}{2!2!}}{\displaystyle _{\mathrm{\Lambda }_4}}u(4,3,2,1)\overline{\psi }(4)\overline{\psi }(3)\psi (2)\psi (1)`$
where $`(i)`$ represents $`(\omega _i,k_i,\mathrm{\Omega }_i)`$ and
$`{\displaystyle _{\mathrm{\Lambda }_4}}:={\displaystyle \underset{i}{}}{\displaystyle \frac{d\mathrm{\Omega }}{(2\pi )^2}_\mathrm{\Lambda }^\mathrm{\Lambda }\frac{dk}{(2\pi )^2}_{\mathrm{}}^{\mathrm{}}\frac{d\omega }{(2\pi )}\mathrm{\Theta }(\mathrm{\Lambda }|๐_4|)}.`$ (24)
with a cutoff function for $`k_4`$, $`\mathrm{\Theta }(\mathrm{\Lambda }|๐_4|)`$, which is needed to make all the momenta lie within the band of width $`2\mathrm{\Lambda }`$ around the Fermi surface. Here $`v_F^{}=k_F/m^{}`$ and
$`m^{}={\displaystyle \frac{1}{Z(1+\frac{m}{k_F}\frac{\mathrm{\Sigma }}{k})}}`$ (25)
is the effective mass of the nucleon which will be equal to the Landau mass $`m_L^{}`$ as will be elaborated on later. $`\mathrm{\Sigma }(\omega ,k)`$ is self-energy and $`1/Z=1+i\frac{\mathrm{\Sigma }}{\omega }`$. The effective mass arises because the eliminated mode contributes to $`i\omega \overline{\psi }\psi `$ and $`k\overline{\psi }\psi `$ differently. By defining $`\psi ^{}=s^{3/2}Z^{1/2}\psi `$ we fix the coefficient of $`i\omega \overline{\psi }\psi `$ and define the effective mass. The term with $`\delta \mu ^{}`$ is a counter term added to assure that the Fermi momentum is fixed (that is, the density is fixed). What this term does is to cancel loop contributions involving the four-Fermi interaction to the nucleon self-energy (i.e., the tadpole) which contributes marginally so that the $`v_F^{}`$ is at the fixed point. Since other contributions which can be written as $`\mathrm{\Sigma }\overline{\psi }\psi `$ are irrelevant, $`m`$ moves to $`m^{}`$ in earlier stage of renormalization but becomes a fixed point characterized by some $`m^{}`$. This means that the counter term essentially assures that the effective mass $`m^{}`$ be at the fixed point. Without this procedure, the term quadratic in the fermion field would be โrelevantโ and hence would be unnatural .
Let us see the quartic coupling $`u`$ at tree level. The cutoff function $`\mathrm{\Theta }(\mathrm{\Lambda }|๐_4|)`$ in (24) makes the coupling depend on angles on the Fermi surface. Since all the momenta are on the thin spherical shell near $`k_F`$, momentum conservation makes the new quartic coupling $`u^{}(\omega ^{},k^{},\mathrm{\Omega }^{})`$ decay after the RG transformation at tree level except for the two cases. One is that $`๐ฒ_3`$ and $`๐ฒ_4`$ are the rotation of $`๐ฒ_1`$ and $`๐ฒ_2`$ around the $`๐ฒ_1+๐ฒ_2`$ as seen in Fig. 2.
In that case, the opening angle $`\mathrm{cos}^1(\widehat{๐}_1\widehat{๐}_2)`$ is fixed where $`\widehat{๐}_i`$ is a unit vector in the direction of $`๐ฒ_i`$. The other, so-called BCS coupling, is that $`\widehat{๐}_1=\widehat{๐}_2`$ and $`\widehat{๐}_3=\widehat{๐}_4`$. These cases are represented by functions;
$`u(\theta _{12}=\theta _{34})`$ $`=`$ $`(\theta _{12},\vartheta )`$ (26)
$`u(\theta _{13}=\theta _{24})`$ $`=`$ $`๐ฑ(\theta _{13})`$
where $`\theta _{ij}=\widehat{๐}_i\widehat{๐}_j`$ and $`\vartheta `$ is the angle between the planes containing $`(๐ฒ_1,๐ฒ_2)`$ and $`(๐ฒ_3,๐ฒ_4)`$ respectively. $``$ and $`๐ฑ`$ are marginal at tree level.
Then the next question is how $``$ and $`๐ฑ`$ evolve at one loop level. Figure 3 shows the one loop diagrams for the evolution.
Integrating out the momentum shell of thickness $`d\mathrm{\Lambda }`$ at $`k=\pm \mathrm{\Lambda }`$ and sending $`\mathrm{\Lambda }/k_F0`$, all the diagrams in Fig. 3 do not contribute to $``$. So $``$ is left marginal to the one loop order. In the case of $`๐ฑ`$, the third diagram in Fig. 3 makes a flow. When we expand $`๐ฑ`$ in terms of angular momentum eigenfunction, $`๐ฑ`$ becomes irrelevant only if all the $`๐ฑ_l`$โs are repulsive. However, if any $`๐ฑ_l`$ is attractive, it becomes relevant and causes BCS instability. We call it BCS channel. Since Landau theory assures that there is no phase transition for one-to-one correspondence between particles and quasiparticles, BCS channel destabilizes in Landau Fermi-liquid theory.
If we divide the angular part of the shell integration in the action (18) into the cells of size $`\mathrm{\Delta }\mathrm{\Omega }\mathrm{\Lambda }/k_F`$, the shell is split into $`k_F/\mathrm{\Lambda }`$ cells. By analogy with large $`N`$ theory each cell corresponds to one species, i.e., $`Nk_F/\mathrm{\Lambda }`$. $`1/N`$ expansion tells that all higher loop corrections except for the bubbles of the first diagram in Fig. 3 are down by powers of $`1/N=\mathrm{\Lambda }/k_F`$. And the contributions of bubbles of the first diagram survive but are irrelevant to the $`\beta `$ function. Only one loop contributions to $`\beta `$ function survive in the large $`N`$ limit. So the four-Fermi interactions in the phonon channel $``$ are also at the fixed points in addition to the Fermi surface fixed point with the effective mass $`m^{}`$. Note that only forward scattering $`(\vartheta =0)`$ is important for responses to soft probes, since nonforward amplitudes in loop calculations are also suppressed by $`1/N`$. Six-Fermi and higher-Fermi interactions are irrelevant and contribute at most screening of the fixed-point constants. The Landau parameter $`F`$ can be identified with forward scattering $`\gamma \frac{m^{}}{2\pi ^2k_F}(\vartheta =0)`$ easily . We arrive at the Fermi-liquid fixed point theory in the absence of BCS interactions.
Chen, Frรถhlich, and Seifert obtain the same result in the $`1/N`$ expansion where their $`N`$ is taken to be $`N\lambda `$ with $`1/\lambda `$ being the width of the effective wave vector space around the Fermi sea which can be considered as the ratio of the microscopic scale to the mesoscopic scale. More specifically if one rescales the four-Fermi interaction such that one defines the dimensionless constant $`gu_0/k_F^2`$ where $`u_0`$ is the leading term (i.e., constant term) in the Taylor series of the quantity $`u`$ in (23), then the fermion wave function renormalization $`Z`$, the Fermi velocity $`v_F`$ and the constant $`g`$ are found not to flow up to order $`๐ช(g^2/N)`$. Thus in the large $`N`$ limit, the system flows to Landau fixed point theory to all orders of loop corrections. This result is correct provided there are no long-range interactions and if the BCS channel is turned off.
### 2.3 Relations for relativistic Fermi-liquid
In this section we briefly summarize the relations of physical properties of the relativistic Landau Fermi-liquid. The extension of Landau Fermi-liquid theory to relativistic region for high density matter is found in . It should be pointed out that one can use all the standard Landau Fermi-liquid relations established below in our calculations once fixed point quantities are identified in the chiral Lagrangian.
#### 2.3.1 Compression modulus and $`F_0`$
The density of states at Fermi surface is
$`N(0)`$ $`=`$ $`\left({\displaystyle \frac{\rho }{\epsilon }}\right)_{k_F}={\displaystyle \frac{\gamma k_Fm_L^{}}{2\pi ^2}}.`$ (27)
Note that $`m_L^{}=\sqrt{m_N^2+k_F^2}`$ for relativistic non-interacting Fermi gas. $`m_L^{}`$ defined by (7) includes the kinetic energy for relativistic Fermi liquid. The chemical potential is defined by
$`\mu {\displaystyle \frac{}{\rho }}=\epsilon _F`$ (28)
where $``$ represents the energy per volume. Using (27) and (28) one can derive the relativistic relation between compression modulus which represents the change of volume with pressure and $`F_0`$ in the same way as the nonrelativistic one:
$`K`$ $`=`$ $`9\rho {\displaystyle \frac{\mu }{\rho }}=9\rho {\displaystyle \frac{}{\rho }}\left(\epsilon _p^0+{\displaystyle ๐\tau ^{}f_{pp^{}}n_p^{}}\right)`$ (29)
$`=`$ $`{\displaystyle \frac{3k_F^2}{m_L^{}}}(1+F_0)`$
Here $`\rho `$ is the baryon number density and $`n_p`$ is the Fermi distribution function for the state of momentum $`p`$.
#### 2.3.2 Landau effective mass
In deriving the Landau mass formula, we shall compare the rest frame and the boosted frame with very small velocity $`๐`$ and check Lorentz symmetry. Since $`u|๐|`$ is small we neglect the order of $`u^2`$ in the process of the derivation. (Note that $`\gamma _u=(1u^2)^{1/2}1`$ in that order. )
Let us first derive the relativistic relation between the Landau effective mass and the velocity dependence of the quasiparticle interaction. When we add a particle (or equivalently quasiparticle) of momentum $`๐`$ in the rest frame to the system, the energy and the momentum of the system increases by $`๐`$ and $`\epsilon _p(0)`$ respectively. In a moving frame with velocity $`๐`$, the momentum and energy increase by
$`๐^{}`$ $`=`$ $`๐\widehat{๐}(\widehat{๐}๐)(1\gamma _u)+\epsilon _p(0)๐\gamma _u`$ (30)
$``$ $`๐+\epsilon _p(0)๐`$
$`\epsilon _p^{}(u)`$ $`=`$ $`(\epsilon _p(0)+๐๐)\gamma _u`$ (31)
$``$ $`\epsilon _p(0)+๐๐.`$
From $`\epsilon _p(u)`$. (30) and (31), we have
$`\epsilon _p(u)=\epsilon _{p\epsilon _p(0)u}(0)+๐๐=\epsilon _p(0)\epsilon _p(0)๐\mathbf{}_p\epsilon _p(0)+๐๐.`$ (32)
and
$`\epsilon _p(u)=\epsilon _p(0)+{\displaystyle ๐\overline{\tau }f_{p\overline{p}}(n_{\overline{p}}(u)n_{\overline{p}}(0))}.`$ (33)
Since $`n_p^{}(u)=n_p(0)`$, we can obtain
$`n_{\overline{p}}(u)n_{\overline{p}\epsilon _{\overline{p}}u}(0)=n_{\overline{p}}(0)e_{\overline{p}}(0)๐\mathbf{}_{\overline{p}}n_{\overline{p}}(0)`$ (34)
using (30) and (31). Then (33) becomes
$`\epsilon _p(u)=\epsilon _p(0){\displaystyle ๐\overline{\tau }f_{p\overline{p}}\epsilon _{\overline{p}}(0)๐\mathbf{}_{\overline{p}}n_{\overline{p}}(0)}.`$ (35)
Comparing it with (32)
$`๐=\epsilon _p(0)\mathbf{}_p\epsilon _p(0){\displaystyle ๐\overline{\tau }f_{p\overline{p}}\epsilon _{\overline{p}}(0)\mathbf{}_{\overline{p}}n_{\overline{p}}(0)}.`$ (36)
In the ground state
$`\mathbf{}_{\overline{p}}n_{\overline{p}}(0)=\delta (\epsilon _{\overline{p}}\mu ){\displaystyle \frac{\epsilon _{\overline{p}}}{\overline{p}}}\widehat{\overline{๐}}`$ (37)
gives
$`p=\epsilon _p{\displaystyle \frac{\epsilon _p}{p}}+{\displaystyle ๐\overline{\tau }\delta (\epsilon _{\overline{p}}\mu )\epsilon _{\overline{p}}f_{p\overline{p}}\widehat{๐}\widehat{\overline{๐}}\frac{\epsilon _{\overline{p}}}{\overline{p}}}.`$ (38)
The chemical potential $`\mu `$ is $`\epsilon _{p_F}`$. Thus (36) becomes on the Fermi surface
$`\mu \left({\displaystyle \frac{\epsilon _p}{p}}\right)_{p_F}=p_F\mu {\displaystyle \frac{\gamma p_F^2}{2\pi ^2}}{\displaystyle \frac{f_1}{3}}.`$ (39)
Using (27), one obtains
$`{\displaystyle \frac{m_L^{}}{\mu }}=1+{\displaystyle \frac{F_1}{3}}.`$ (40)
This is the relativistically extended relation of the famous Landau mass formula
$$\frac{m_L^{}}{M}=1+\frac{F_1}{3},$$
(41)
that follows from Galilean invariance in the same way .
#### 2.3.3 First sound velocity
The first sound is a density oscillation mode under the circumstances where there are many quasiparticle collisions during the time of interest. The sufficiently often collisions produce the required local equilibrium in the time scale of the period of motion.
When first sound wave gives small change in density of a static homogeneous relativistic fluid in a comoving frame without changing entropy, the continuity equation requires
$`{\displaystyle \frac{}{t}}\delta \rho =\rho \mathbf{}๐.`$ (42)
Under the condition of the fixed entropy
$`kds={\displaystyle \frac{1}{T}}(p\delta v+\delta \epsilon )={\displaystyle \frac{1}{\rho ^2T}}\{\rho \delta \epsilon (p+)\delta \rho \}=0`$ (43)
with the entropy per particle $`ks`$, the volume per particle $`v=1/\rho `$, and the energy per particle $`\epsilon =/\rho `$, the relatvistic equation of motion becomes
$`{\displaystyle \frac{๐}{t}}={\displaystyle \frac{1}{p+}}\mathbf{}\delta p={\displaystyle \frac{\delta p}{\rho \delta }}\mathbf{}\delta \rho .`$ (44)
Applying (44) to (42), we obtain
$`{\displaystyle \frac{^2}{t^2}}\delta \rho =\left({\displaystyle \frac{p}{}}\right)\mathbf{}^2\delta \rho .`$ (45)
Thus the first sound velocity of the relativistic Landau Fermi-liquid is
$`c_1^2`$ $``$ $`{\displaystyle \frac{p}{}}`$
$`=`$ $`{\displaystyle \frac{\rho }{}}{\displaystyle \frac{}{\rho }}\left(\rho ^2{\displaystyle \frac{/\rho }{\rho }}\right)`$
$`=`$ $`{\displaystyle \frac{K}{9\mu }}`$
$`=`$ $`{\displaystyle \frac{k_F^2}{3\mu ^2}}{\displaystyle \frac{1+F_0}{1+F_1/3}}`$
from the above results.
## 3 Chiral effective Lagrangian for nuclei
In this section we come to hadron/nuclear phenomena. We simply explain the basic theory of strong interaction, QCD, and how successfully low energy phenomena can be described by its effective theory. Then it is discussed how the effective theory can be applied to the nuclear/hadronic matter ground state. FTS1 model will be studied as an example.
### 3.1 QCD: basis of strong interactions
Quantum Chromodynamics is believed to be the fundamental theory for the strong interaction. It is a nonabelian gauge description of the strong interaction and its building blocks are quarks and gluons. The QCD Lagrangian is
$`_{QCD}={\displaystyle \frac{1}{2}}\mathrm{Tr}G_{\mu \nu }G^{\mu \nu }+\overline{q}(i/gG/m_q)q`$ (47)
with $`q=(u,d,\mathrm{})^T`$ and $`m_q=`$diag$`[m_u,m_d,\mathrm{}]`$ in flavor space. In this section $`\mathrm{Tr}`$ represents the trace over flavor indices. The gauge field strength tensor is
$`G_{\mu \nu }=_\mu G_\nu _\nu G_\mu ig[G_\mu ,G_\nu ]`$ (48)
where $`G_\mu =G_\mu ^a\frac{\lambda ^a}{2}`$. $`\lambda ^a`$ is the well-known Gell-Mannโs $`3\times 3`$ traceless hermitian matrix with color indices $`a=1,2,\mathrm{},8`$.
Quantum Chromodynamics conserves its relevant symmetries; Lorentz symmetry, parity, charge conjugation, time reversal and color SU(3) gauge symmetry. In addition, it has approximate symmetries. Isospin symmetry is a good one because of small difference between $`u`$ and $`d`$ masses.
There is another well-known and important QCD approximate symmetry, chiral symmetry. It is based on the fact that the mass of light quarks $`u`$ and $`d`$ are much smaller than chiral symmetry breaking scale $`\mathrm{\Lambda }_\chi 4\pi f_\pi 1`$ GeV. We can consider the massless limit, i.e. $`m_q0`$ when dealing with light quark physics. It means that chiral symmetry becomes relevant. With projection operator
$`P_\pm ={\displaystyle \frac{1\gamma _5}{2}}`$ (49)
we can decompose the spinor into the eigenstates of helicity $`๐\widehat{๐}`$,
$`๐\widehat{๐}q=\pm q_\pm .`$ (50)
with $`q_\pm P_\pm q`$. Without quark mass term, $`q_+`$ and $`q_{}`$ are decoupled each other
$`_{QCD}{\displaystyle \frac{1}{2}}\mathrm{Tr}G_{\mu \nu }G^{\mu \nu }+\overline{q}_+(i/gG/)q_+\overline{q}_{}(i/gG/)q_{}`$ (51)
So the separate transformations
$`q_\pm e^{iฯต_\pm }q_\pm `$ (52)
and
$`q_\pm e^{i๐ฝ_\pm ๐}q_\pm `$ (53)
with real parameters $`ฯต_\pm `$ and $`๐ฝ_pm`$ leave the massless QCD Lagrangian invariant, where $`๐`$ is the general SU(2) generators. We can construct SU(2) and U(1) vector transformations as
$`qe^{i๐ฝ_V๐}q,qe^{iฯต_V}q`$ (54)
and axial ones as
$`qe^{i๐ฝ_A๐\gamma _5}q,qe^{iฯต_A\gamma _5}q`$ (55)
with $`๐ฝ_\pm =๐ฝ_V\pm ๐ฝ_A`$ and $`ฯต_\pm =ฯต_V\pm ฯต_A`$. The corresponding currents are
$`๐ฝ^\mu `$ $`=`$ $`\overline{q}\gamma ^\mu ๐q,V_s^\mu =\overline{q}\gamma ^\mu q,`$ (56)
$`๐จ^\mu `$ $`=`$ $`\overline{q}\gamma ^\mu \gamma _5๐q,A_s^\mu =\overline{q}\gamma ^\mu \gamma _5q.`$ (57)
Although massless QCD Lagrangian is invariant under these transformations, its U(1) axial current of QCD is not conserved due to anomaly even if the quark mass $`m_q0`$;
$`_\mu A_s^\mu ={\displaystyle \frac{N_Fg^2}{16\pi ^2}}\mathrm{Tr}G_{\mu \nu }\stackrel{~}{G}^{\mu \nu }+2i\overline{q}m_q\gamma _5q`$ (58)
with $`\stackrel{~}{G}^{\mu \nu }=\epsilon ^{\mu \nu \alpha \beta }G_{\alpha \beta }`$. So the massless QCD has chiral symmetry (SU(2)$`{}_{R}{}^{}\times `$SU(2)<sub>L</sub>) and fermion number symmetry (U(1)<sub>V</sub>). Chiral symmetry does not appear in hadron spectrum. So it is generally assumed that the chiral symmetry is dynamically broken into SU(2)<sub>V</sub> in hadron physics and the light pseudoscalar mesons are regarded as Goldstone bosons.
Since the massless QCD Lagrangian contains no dimensional parameters, massless QCD action is invariant under scale transformation;
$`x^{}=ax,q(x^{})=a^{3/2}q(x),G_\mu (x^{})=a^1G_\mu (x).`$ (59)
Thus massless QCD seems to have another approximate symmetry, scale symmetry. However, the renormalization prescriptions which have the running coupling constant break the scale symmetry. It means that we need to introduce a dimensional scale in order to specify the value of running coupling constant. Such specification of a scale breaks the scale invariance of QCD;
$`_\mu D^\mu =\theta _\mu ^\mu ={\displaystyle \frac{\beta }{g}}\mathrm{Tr}G_{\mu \nu }G^{\mu \nu }+\overline{q}(1+\gamma _q)m_qq`$ (60)
with the anomalous dimension of quarks $`\gamma _q=`$diag$`[\gamma _u,\gamma _d,\mathrm{}]`$ even if quark masses are zero. $`D^\mu `$ is dilatation current, $`\theta _\nu ^\mu `$ is the improved energy-momentum tensor and
$`\beta \mu {\displaystyle \frac{g}{\mu }}=({\displaystyle \frac{2}{3}}N_F11){\displaystyle \frac{g^3}{16\pi ^2}}+๐ช(g^5)`$ (61)
with the scale of renormalization $`\mu `$. Contrary to the axial anomaly which includes only one-loop contributions, scale anomaly includes multi-loop contributions in $`\beta `$. Scale anomaly which plays a major role in breaking scale invariance of light quark QCD gives a basis of BR scaling argument which will appear in Section 4.2.
### 3.2 Effective field theories
We know that the particles which appear in nuclear physics, e. g. pions, nucleons, etc., are not elementary particles. In the framework of QCD, they are known to consist of quarks and gluons which give non-perturbative contributions in low energy processes. In addition, QCD is a part of the Standard Model which is also an effective theory, not a fundamental theory.
In order to deal with low energy hadrons, one can construct an effective field theory that is appropriate to the probed energy scale $`Q`$. The relevant degrees of freedom in such effective field theory are the low-energy particles which appear at the energy scale of the observed experiments. The particles with energies higher than the relevant energy scale are integrated out and absorbed into the couplings among the relevant degrees of freedom;
$`{\displaystyle [๐q][๐\overline{q}][๐g]e^{i{\scriptscriptstyle d^4x_{fund}[q,\overline{q},g,\eta ,\overline{\eta },j]}}}`$ (62)
$`=`$ $`{\displaystyle [๐B][๐\overline{B}][๐M]e^{i{\scriptscriptstyle d^4x_{eff}[B,\overline{B},M,\eta ,\overline{\eta },j]}}}`$
where $`\eta `$, $`\overline{\eta }`$, and $`j`$ are the external sources of elementary fermions $`q`$(quarks), anti-fermions $`\overline{q}`$(anti-quarks), and bosons $`g`$(gluons) respectively in fundamental theory(QCD) and $`B`$, $`\overline{B}`$, and $`M`$ represent fermions(baryons), anti-fermions(anti-baryons), and bosons(mesons) respectively in effective field theory. Although one can in principle calculate the couplings in effective theory for strong interactions from the fundamental QCD from equation (62), it seems an impossible mission because we do not know how to perform such a highly non-perturbative calculation. However, we do not have to deal with, nor to know exactly, the fundamental theory.
The strategy to build an effective theory is simple. It consists of writing down the most general Lagrangian which conserves all the relevant symmetries that figure at a given energy scale, and satisfy the basic principles(e.g. quantum mechanics, cluster decomposition) of the theory. Then any theory under the same constraints looks like his/hers at sufficiently low energy scale though he/she cannot insist that the right theory necessarily leads only to his/hers. (This is called โfolk theoremโ by Weinberg though he used it to explain the usefulness of quantum field theory.) By imposing the relevant symmetries of the QCD or of a more fundamental theory, e.g., chiral symmetry, and the basic principles we can build an effective field theory for low-energy nuclear processes where the composite hadrons are taken as elementary.
The number of terms in the effective Lagrangian which are consistent with fundamental or assumed symmetries may be infinite, but one can manage to describe the probed physical processes by expanding the terms in $`Q/\mathrm{\Lambda }`$, where $`\mathrm{\Lambda }`$ is s suitable cut-off scale and $`Q`$ the scale probed ($`Q<<\mathrm{\Lambda }`$). So only a few leading terms are relevant for low-energy processes. And the couplings can be determined from available experiments.
Such effective field theories with chiral symmetry so designed work very well in matter-free or dilute space. They are used to calculate soft-pion processes firstly , and recently extended to the process in light nuclei though nucleons are not soft . Park et al.โs work for two-nucleon systems at very low energies is one of the best and newest examples of the success. Setting the cutoff near one pion mass, they integrate out all mesonic degrees of freedom, even the pions. Since the deuteron bound state is dilute, the parameters of the theory can be determined by free space experiments. The results in confirm that the strategy of the effective field theory works remarkably well. When the pion field is included in addition, it provides a new degree of freedom and improves the theory even further allowing one to go higher in energy scale .
But in heavier nuclei, the energy scales of the system will be higher since the interactions between nucleons in such systems sample all length scales and hence other degrees of freedom than nucleonic and pionic need be introduced. It means that the irrelevant and neglected terms in the computation for the light nuclei become more relevant. We need to consider more and more terms as the density of the system becomes higher and higher. The power of effective field theory whose parameters are determined from the matter-free experiments gets weakened in dense system. How can we proceed as density increases beyond the ordinary matter density for which there are practically no experimental data?
### 3.3 Effective Lagrangian in medium and Landau theory
Recently Lynn made a progress to give a good hint to answer the above questions . He proposed that the ground-state matter is โchiral liquidโ which arises as a non-topological soliton. Fluctuation around this ground state should give an accurate description of the observables that we are dealing with. We shall here extend this argument further and make contact with Landauโs Fermi-liquid theory of nuclear matter. This will allow us to understand the nuclear/hadron matter description in terms of chiral Lagrangians and Fermi-liquid fixed point theory thereby giving a unified picture of ordinary nuclear matter and extreme state of matter probed in heavy-ion collisions. This is the attempt to connect the physics of the two vastly different regimes.
The basic assumption we start with is that the chiral liquid arises from a quantum effective action resulting from integrating out the degrees of freedom lying above the chiral scale $`\mathrm{\Lambda }_\chi 4\pi f_\pi 1`$ GeV.
$`{\displaystyle [d\varphi _<]e^{iS_\chi (\varphi _<)}}={\displaystyle [d\varphi _<][d\varphi _>]e^{iS(\varphi _<,\varphi _>)}}`$ (63)
where the subscript $`<`$($`>`$) represents the sector $`\omega <\mathrm{\Lambda }_\chi `$($`\omega >\mathrm{\Lambda }_\chi `$) of the given set of fields $`\varphi `$. As explained in Sec. 3.2,
$`S_\chi ={\displaystyle \underset{i}{}}g_i\widehat{O}_i`$ (64)
is the sum of all possible terms consistent with symmetries of QCD. This corresponds to the first stage of โdecimationโ in our scheme. The mean field solution of this action is then supposed to yield the ground state of nuclear matter with the Fermi surface characterized by the Fermi momentum $`k_F`$. The effective Lagrangian was given in terms of the baryon, pion, quarkonium scalar and vector fields. The gluonium scalars are integrated out. Instead of treating the scalar and vector fields explicitly, we will consider here integrating them out further from the effective Lagrangian. This will lead to four-Fermi, six-Fermi, etc. , interactions in the Lagrangian with various powers of derivatives acting on the Fermi field. The resulting effective Lagrangian will then consist of the baryons and pions coupled bilinearly in the baryon field and four-Fermi and higher-Fermi interactions with various powers of derivatives, all consistent with chiral symmetry. A minimum version of such Lagrangian in mean field can be shown to lead to the original (naive) Walecka model .
The next step is to decimate successively the degrees of freedom present in the excitations with the scale $`E<\mathrm{\Lambda }_\chi `$ . To do this, we consider excitations near the Fermi surface, which we take to be spherical for convenience characterized by $`k_F`$. First we integrate out the excitations with momentum $`p\pm \mathrm{\Lambda }`$ (where $`p=|๐|`$ and $`\mathrm{\Lambda }<\mathrm{\Lambda }_\chi `$) measured relative to the Fermi surface (corresponding to the particle-hole excitations with momentum greater than $`2\mathrm{\Lambda }`$). We are thus restricting ourselves to the physics of excitations whose momenta lie below $`2\mathrm{\Lambda }`$ as in Section 2.2. Leaving out the pion for the moment and formulated non-relativistically, the appropriate action to consider can be written in a simplified and schematic form as Eq. (23).<sup>1</sup><sup>1</sup>1The pion will be introduced in the Section 6.1 in terms of a non-local four-Fermi interaction that enters in the ground state property and gives the nucleon Landau mass formula in terms of BR scaling and pionic Fock term.
In nuclear matter, the spin and isospin degrees of freedom need to be taken into account into the four-Fermi interaction. All these can be written symbolically in the action (23). The function $`u`$ in the four-Fermi interaction term therefore contains spin and isospin factors as well as space dependence that takes into account non-locality and derivatives. For simplicity we will consider it to be a constant depending in general on spin and isospin factors. Non-constant terms will be โirrelevant.โ In our discussions, the BCS channel that corresponds to a particle-particle channel does not figure and hence will not be considered explicitly.
The successive mode elimination
$`{\displaystyle [d\varphi _<^l]e^{iS^{}(\varphi _<^l)}}={\displaystyle [d\varphi _<^l][d\varphi _<^h]e^{iS_\chi (\varphi _<^l,d\varphi _<^h)}},`$ (65)
which satisfies RG equation explained in Sec. 2.2, will give
$`S^{}={\displaystyle \underset{i}{}}g_i^{}\widehat{O}_i^{}.`$ (66)
The starred coupling constant $`g_i^{}`$ and operators $`\widehat{O}_i^{}`$ depend on density through $`s`$ and have the same structure of (64). $`l(h)`$ represents the components $`p<\mathrm{\Lambda }/s`$($`p>\mathrm{\Lambda }/s`$). The upshot of the analysis in Section 2.2 is that the resulting theory is the Fermi-liquid fixed point theory with the limit $`\mathrm{\Lambda }/k_F0`$. In sum, we arrive at the picture where the chiral liquid solution of the quantum effective chiral action gives the Fermi-liquid fixed point theory. The parameters of the four-Fermi interactions in the phonon channel are then identified with the fixed-point Landau parameters.
There are two steps to apply such scheme to nuclear/hadron phenomenology. The first is to derive the in-medium effective Lagrangian directly from QCD or from the matter-free effective Lagrangian and the second is to solve the built in-medium effective Lagranian. The first is very difficult. So we usually go to the second step, after building the in-medium effective Lagrangian by reasonable guesses. We assume that the effective Lagrangian satisfies
$`S^{eff}={\displaystyle d^4x^{eff}}`$ (67)
where $`S^{eff}`$ is a Wilsonian effective action arrived at after integrating out high-frequency modes of the nucleon and other heavy degrees of freedom. This action is then given in terms of sum of terms organized in chiral order in the sense of effective theories at low energy.
One way to build the chiral effective Lagrangian for nuclear matter has been studied by Furnstahl, Serot and Tang <sup>2</sup><sup>2</sup>2The model in is referred to as FTS1. That in shall as FTS2. They formulated their theory in terms of a chiral Lagrangian constructed by using the terms which are governed by QCD symmetry and applying the โnaturalnessโ condition for all relevant fields. In doing this, they introduced in the FTS1 a quarkonium field that is associated with the trace anomaly with its potential constrained by Vainshtein et al.โs low energy theorem . And Georgiโs โnaive dimensional analysisโ <sup>3</sup><sup>3</sup>3Since the strong interactions have two relevant scales $`f_\pi `$ and $`\mathrm{\Lambda }_\chi `$ and $`\mathrm{\Lambda }_\chi `$ is much larger than $`f_\pi `$, we can apply Georgiโs naive dimensional analysis to the low energy hadron physics. was used in the FTS2 instead of the trace anomaly. It was argued in that a Lagrangian so constructed contains in principle arbitrarily higher-order many-body effects including those loop corrections that can be expressed as counter terms involving matter fields (e.g., baryons). This is essentially equivalent to Lynnโs chiral effective action that purports to include all orders of quantum loops in chiral expansion supplemented with counter terms consistent with the order to which loops are calculated. Though it is a little hard to define the fluctuation on its ground state, their models are very successful in describing the bulk properties of nuclei.
Another way is to apply the strategy of the effective theory to the in-medium Lagrangian. The parameters of the effective Lagrangian are related to the vacuum state at a given density so depend on the density. One famous example is BR scaling . Brown and Rho point that the mean field solution of the chiral effective Lagrangian given by BR scaling approximates
$`\delta S^{eff}=0.`$ (68)
The detail of BR scaling is reviewed in the next section. The aim of this review is to cast BR scaling in a suitable form starting with a chiral Lagrangian description of the ground state as specified above around which fluctuations in a various flavour sectors are to be made. To do this we study phenomenologically successful FTS1 model here.
### 3.4 A chiral effective model: FTS1 model
Furnstahl, Tang and Serot constructed an effective nonlinear chiral model that will be referred to as FTS1 model in this review. The FTS1 model incorporates the scale anomaly of QCD in terms of a light (โquarkoniumโ) scalar field $`S`$ and a heavy (โgluoniumโ) scalar field $`\varsigma `$ and gives a good description of basic nuclear properties in mean field. One can also build a model appealing to general notions of effective field theories such as โnaturalness conditionโ as in This avoids the use of the scale anomaly of QCD. The FTS2 model is also an effective mean field theory which gives an equally satisfactory phenomenology as the FTS1. However the FTS1 is found to be more convenient for studying the role of the light scalar field in the scale anomaly we are interested in since the FTS1 Lagrangian includes the scale anomaly term explicitly. In addition, Li, Brown, Lee, and Ko found that the FTS1 model reproduces quite successfully nucleon flow in heavy ion collisions .
The underlying assumption in the FTS1 is that the light scalar field transforms under scale transformation as
$`S(a^1x)=a^dS(x)`$ (69)
with a parameter $`d`$ that can be different from its canonical scale dimension, i.e. unity, while the scale dimension of the heavy gluonium, which is integrated out in the effective Lagrangian for normal nuclear matter, is taken to be unity. This assumption imposes that quantum fluctuations in the scalar channel be incorporated into an anomalous dimension $`d_{an}=d10`$. A RG flow argument in Section 3.3 justifies this assumption heuristically. One further assumption of the FTS1 is that there is no mixing between the light scalar $`S(x)`$ and the heavy scalar $`\varsigma `$ in the trace anomaly. The FTS1 Lagrangian has the form
$`^{eff}=_sH_g{\displaystyle \frac{\varsigma ^4}{\varsigma _0^4}}\left(\mathrm{ln}{\displaystyle \frac{\varsigma }{\varsigma _0}}{\displaystyle \frac{1}{4}}\right)H_q\left({\displaystyle \frac{S^2}{S_0^2}}\right)^{\frac{2}{d}}\left({\displaystyle \frac{1}{2d}}\mathrm{ln}{\displaystyle \frac{S^2}{S_0^2}}{\displaystyle \frac{1}{4}}\right)`$ (70)
where $`_s`$ is the chiral- and scale-invariant Lagrangian containing $`\varsigma ,S,N,\pi ,\omega `$, etc. Here $`\varsigma _0`$ and $`S_0`$ are the vacuum expectation values(VEV) with the vacuum $`|0`$ defined in the matter-free space:
$`\varsigma _00|\varsigma |0,S_00|S|0.`$ (71)
The trace of the improved energy-momentum tensor , i. e. the divergence of the dilatation current $`D^\mu `$, from the Lagrangian is;
$`_\mu D^\mu =\theta _\mu ^\mu =H_g{\displaystyle \frac{\varsigma ^4}{\varsigma _0^4}}H_q\left({\displaystyle \frac{S^2}{S_0^2}}\right)^{2/d}.`$ (72)
The mass scale associated with the gluonium degree of freedom is higher than that of chiral symmetry, $`\mathrm{\Lambda }_\chi 1`$ GeV. For example the mass of the scalar glueball calculated by Weingarten is 1.6$``$1.8 GeV. This invites us to integrate out the gluonium. The resulting FTS1 effective Lagrangian takes the form
$``$ $`=`$ $`\overline{N}[i\gamma _\mu (^\mu +iv^\mu +ig_v\omega ^\mu +g_A\gamma _5a^\mu )M+g_s\varphi ]N`$ (73)
$`{\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }+{\displaystyle \frac{1}{4!}}\zeta g_v^4(\omega _\mu \omega ^\mu )^2`$
$`+{\displaystyle \frac{1}{2}}\left(1+\eta {\displaystyle \frac{\varphi }{S_0}}\right)\left\{{\displaystyle \frac{f_\pi ^2}{2}}\mathrm{tr}(_\mu U^\mu U^{})+m_v^2\omega _\mu \omega ^\mu \right\}`$
$`+{\displaystyle \frac{1}{2}}_\mu \varphi ^\mu \varphi {\displaystyle \frac{m_s^2}{4}}S_0^2d^2\left[\left(1{\displaystyle \frac{\varphi }{S_0}}\right)^{4/d}\left\{{\displaystyle \frac{1}{d}}\mathrm{ln}\left(1{\displaystyle \frac{\varphi }{S_0}}\right){\displaystyle \frac{1}{4}}\right\}+{\displaystyle \frac{1}{4}}\right]`$
where $`S=S_0\varphi `$, $`\eta `$ and $`\zeta `$ are unknown parameters to be fixed and
$`\xi ^2`$ $`=`$ $`U=e^{i๐
๐/f_\pi }`$
$`v_\mu `$ $`=`$ $`{\displaystyle \frac{i}{2}}(\xi ^{}_\mu \xi +\xi _\mu \xi ^{})`$
$`a_\mu `$ $`=`$ $`{\displaystyle \frac{i}{2}}(\xi ^{}_\mu \xi \xi _\mu \xi ^{}).`$
Note that a given VEV of the $`\varphi `$ field scales down the pion decay constant and the $`\omega `$ mass in the same way at the lowest chiral order as in . The static mean field equations of motion for FTS1 are
$`g_s{\displaystyle \underset{i}{}}\overline{N}_iN_i`$ $`=`$ $`^2\varphi _0m_s^2\left(1{\displaystyle \frac{\varphi _0}{S_0}}\right)^{\frac{4d}{d}}\mathrm{ln}\left(1{\displaystyle \frac{\varphi _0}{S_0}}\right){\displaystyle \frac{\eta }{2S_0}}m_v^2\omega _0^2`$ (74)
$`g_v{\displaystyle \underset{i}{}}\overline{N}_i^{}N_i`$ $`=`$ $`^2\omega _0+m_v^2\left(1+\eta {\displaystyle \frac{\varphi _0}{S_0}}\right)\omega _0+{\displaystyle \frac{\zeta }{6}}g_v^4\omega _0^3`$ (75)
with the static mean field solutions $`\varphi _0`$ and $`\omega _0`$. Equation (75) is a constraint because $`\omega _0`$ is not a dynamical degree of freedom.
It is important to note that the FTS1 Lagrangian is an effective Lagrangian in the sense explained in Section 3.2. The effect of high frequency modes of the nucleon field and other massive degrees of freedom appears in the parameters of the Lagrangian and in the counter terms that render the expansion meaningful. It presumably includes also vacuum fluctuations in the Dirac sea of the nucleons . In general, it could be much more complicated. Indeed, one does not yet know how to implement this strategy in full rigor given that one does not know what the matching conditions are. In , the major work is, however, done by choosing the relevant parameters of the FTS1 Lagrangian to fit the empirical informations.
The energy density for uniform nuclear matter obtained from (73) is
$`\epsilon `$ $`=`$ $`{\displaystyle \frac{\gamma }{(2\pi )^3}}{\displaystyle ^{k_F}}d^3k\sqrt{๐^2+(Mg_s\varphi _0)^2}`$ (76)
$`{\displaystyle \frac{m_v^2}{2}}\left(1+\eta {\displaystyle \frac{\varphi _0}{S_0}}\right)\omega _0^2+g_v\rho _B\omega _0{\displaystyle \frac{\zeta }{4!}}g_v^4\omega _0^4`$
$`+{\displaystyle \frac{m_s^2}{4}}S_0^2d^2\left\{\left(1{\displaystyle \frac{\varphi _0}{S_0}}\right)^{\frac{4}{d}}\left({\displaystyle \frac{1}{d}}\mathrm{ln}(1{\displaystyle \frac{\varphi _0}{S_0}}){\displaystyle \frac{1}{4}}\right)+{\displaystyle \frac{1}{4}}\right\}.`$
Here $`\gamma `$ is the degeneracy factor.
### 3.5 Anomalous dimension in FTS1 model
The best fit to the properties of nuclear matter and finite nuclei is obtained with the parameter set T1<sup>4</sup><sup>4</sup>4Explicitly the T1 parameters are: $`g_s^2=99.3`$, $`g_v^2=154.5`$, $`\eta =0.496`$, and $`\zeta =0.0402`$. when the scale dimension of the scalar $`S`$ is near $`d=2.7`$ in the FTS1 . The large anomalous dimension means that one is fluctuating around a wrong ground state. Brown-Rho scaling is meant to avoid this. In this section, we analyze how this comes out and present what we understand on the role of the large anomalous dimension $`d_{an}=d11.7`$ in nuclear dynamics. In what follows, the parameter T1 with this anomalous dimension will be taken as a canonical parameter set.
#### 3.5.1 Scale anomaly
Following Coleman and Jackiw , the scale anomaly can be discussed in terms of an anomalous Ward identity. Define $`\mathrm{\Gamma }_{\mu \nu }(p,q)`$ and $`\mathrm{\Gamma }(p,q)`$ by
$`G(p)\mathrm{\Gamma }_{\mu \nu }(p,q)G(p+q)`$ $`=`$ $`{\displaystyle d^4xd^4ye^{iqx}e^{ipy}0T^{}\theta _{\mu \nu }(x)\phi (y)\phi (0)0}`$ (77)
$`G(p)\mathrm{\Gamma }G(p+q)`$ $`=`$ $`{\displaystyle d^4xd^4ye^{iqx}e^{ipy}0T^{}[^\mu D_\mu ](x)\phi (y)\phi (0)0}`$
with the renormalized propagator $`G(p)`$ and the renormalized fields $`\phi (x)`$. Here $`T^{}`$ represents the covariant T-product, $`D_\mu (x)`$ the dilatation current, and $`\theta _{\mu \nu }`$ the improved energy momentum tensor with $`D_\mu =x^\nu \theta _{\mu \nu }`$. A naive consideration on Ward identities would give
$`g_{\mu \nu }\mathrm{\Gamma }^{\mu \nu }(p,q)=\mathrm{\Gamma }(p,q)idG^1(p)idG^1(p+q)`$ (78)
with $`d`$ the scale dimension of $`\phi (x)`$. However $`\mathrm{\Gamma }`$ is ill-defined due to singularity and so has to be regularized. With the regularization, the Ward identity reads
$`g_{\mu \nu }\mathrm{\Gamma }^{\mu \nu }(p,q)`$ $`=`$ $`\mathrm{\Gamma }(p,q)idG^1(p)idG^1(p+q)+A(p,q)`$ (79)
$`A(p,q)`$ $``$ $`\underset{\mathrm{\Lambda }\mathrm{}}{lim}\mathrm{\Gamma }(p,q,\mathrm{\Lambda })\mathrm{\Gamma }(p,q)`$ (80)
where the additional term, $`A`$, is the anomaly. This anomaly corresponds to a shift in the dimension of the field involved at the lowest loop order but at higher orders there are vertex corrections. One obtains however a simple result when the beta functions vanish at zero momentum transfer . Indeed in this case, the only effect of the anomaly will appear as an anomalous dimension. In general this simplification does not occur. However it can take place when there are nontrivial fixed points in the low energy theory. Under the reasoning developed in condensed matter physics it is argued later that nuclear matter is given, in the absence of BCS channel, by a Landau Fermi-liquid fixed point theory with vanishing beta functions of the four-Fermi interactions and that all quantum fluctuation effects would therefore appear in the anomalous dimension of the scalar field $`S`$. That nuclear matter is a Fermi-liquid fixed point seems to be well verified at least phenomenologically as seen in Sect. 5. However that fluctuations into the scalar channel can be summarized into an anomalous dimension is a conjecture that remains to be proved. We conjecture here that this is one way we can understand the success of the FTS1 model.
#### 3.5.2 Nuclear matter properties at $`d_{an}5/3`$
The FTS1 theory has some remarkable features associated with the large anomalous dimension. Particularly striking is the dependence on the anomalous dimension of the compression modulus $`K`$ and many-body forces.
In Table 1 are listed the compression modulus $`K`$ and the equilibrium Fermi momentum $`k_{eq}`$ vs. the scale dimension $`d`$ of the scalar field $`\varphi `$. As the $`d`$ increases, the $`K`$ drops very rapidly and stabilizes at $`K200`$ MeV consistent with experiments for $`d2.6`$ and stays nearly constant for $`d>2.6`$. This can be seen in Fig. 4. The equilibrium Fermi momentum on the other hand slowly decreases uniformly as the $`d`$ increases.
Unfortunately, we have no simple understanding of the mechanism that makes the compression modulus $`K`$ stabilize at the particular value $`d_{an}5/3`$. We believe there is a non-trivial correlation between this behavior of $`K`$ and the observation made below that the scalar logarithmic interaction brought in by the scale anomaly is entirely given at the saturation point by the quadratic term at the same $`d_{an}`$ with the higher polynomial terms (i.e. many-body interactions) contributing more repulsion for increasing anomalous dimension. At present our understanding is purely numerical and hence incomplete.
In mean field, the logarithmic potential in Eqs. (73) and (76) contains n-body-force (for $`n2`$) contributions to the energy density. For the FTS1 parameters, these n-body terms are strongly suppressed for $`d>2.6`$. This is shown in Fig. 5 where it is seen that the entire potential term is accurately reproduced by the quadratic term $`\frac{1}{2}m_s^2\varphi ^2`$ for $`d_{an}5/3`$. Furthermore a close examination of the results reveals that each of the n-body terms are separately suppressed. This phenomenon is in some sense consistent with chiral symmetry and is observed in the spectroscopy of light nuclei .
Since there are additional polynomial terms in vector fields (i.e. terms like $`\varphi \omega ^2`$), the near complete suppression of the scalar polynomials does not mean the same for the total many-body forces. In fact we should not expect it. To explain why this is so, we plot in Fig. 6 the three-body contributions of the $`\varphi ^3`$ and $`\varphi \omega ^2`$ forms. We also compare the FTS1 results with the FTS2 results that are based on the naturalness condition. In FTS1 the $`\varphi ^3`$ term that turns to repulsion from attraction for $`d>8/3`$ contributes little, so the main repulsion arises from the $`\varphi \omega ^2`$-type term. This, together with an attraction from a $`\omega ^4`$ term, is needed for the saturation of nuclear matter at the right density. This raises the question as to how one can understand the result obtained by Brown, Buballa, and Rho where it is argued that the chiral phase transition in dense medium is of mean field with the bag constant given entirely by the quadratic term $`\frac{1}{2}m_s^2\varphi ^2`$. The answer to this question is as follows. First we expect that the anomalous dimension will stay locked at $`d_{an}=d15/3`$ near the phase transition (this is because the anomalous dimension associated with the trace anomaly โ a consequence of ultraviolet regularization โ is not expected to depend upon density), so the $`\varphi ^n`$ terms for $`n>2`$ will continue to be suppressed as density approaches the critical value. Secondly near the chiral phase transition the gauge coupling of the vector meson will go to zero in accordance with the Georgi vector limit , where chiral symmetry is restored by $`m_\rho 0`$, and vector meson decoupling takes place as argued in . So the many-body forces associated with the vector mesons will also be suppressed as density increases to the critical density.
#### 3.5.3 Anomalous dimension and the scalar meson mass
We would like to understand how the large anomalous dimension needed here could arise in the theory and its role in the scalar sector. As suggested in and elaborated more in Section 3.3, one appealing way of understanding the FTS1 mean field theory is to consider all channels to be at Fermi-liquid fixed points except that because of scale anomaly, the scalar field develops an anomalous dimension, thereby affecting four-Fermi interaction in the scalar channel resulting when the scalar field is integrated out. If the anomalous dimension were sufficiently negative so that marginal terms became marginally relevant, then the system would become unstable as in the case of the NJL model or superconductivity, with the resulting spontaneous symmetry breaking. However if the anomalous dimension is positive, then the resulting effect will instead be a screening. The positive anomalous dimension we need here belongs to the latter case. We can see this as follows. Consider the potential given with the low-lying scalar $`S`$ (with the gluonium component integrated out):
$`V(S,\mathrm{})={\displaystyle \frac{1}{4}}m_S^2d^2S_0^2\left({\displaystyle \frac{S}{S_0}}\right)^{\frac{2}{d}}\left({\displaystyle \frac{1}{d}}\mathrm{ln}{\displaystyle \frac{S}{S_0}}{\displaystyle \frac{1}{4}}\right)+\mathrm{}`$ (81)
where $`m_S`$ is the light-quarkonium mass in the free space ($`700`$ MeV) and the ellipses stand for other contributions such as pions, quark masses etc. that we are not interested in. The scalar excitation on a given background $`S^{}`$ is given by the double derivative of $`V`$ with respect to $`S`$ at $`S=S^{}`$
$`m_{S}^{}{}_{}{}^{2}=m_S^2\left({\displaystyle \frac{S^{}}{S_0}}\right)^{\frac{4}{d}2}\left[1+\left({\displaystyle \frac{4}{d}}1\right)\mathrm{ln}{\displaystyle \frac{S^{}}{S_0}}\right].`$ (82)
We may identify the ratio $`S^{}/S_0`$ with the BR scaling factor $`\mathrm{\Phi }`$ :
$`{\displaystyle \frac{S^{}}{S_0}}=\mathrm{\Phi }={\displaystyle \frac{f_\pi ^{}}{f_\pi }}={\displaystyle \frac{m_v^{}}{m_v}}.`$ (83)
Then we have
$`{\displaystyle \frac{m_S^{}}{m_S}}=\mathrm{\Phi }(\rho )\kappa _d(\rho )`$ (84)
with
$`\kappa _d(\rho )=\mathrm{\Phi }^{\frac{2}{d}2}\left[1+\left({\displaystyle \frac{4}{d}}1\right)\mathrm{ln}\mathrm{\Phi }\right]^{\frac{1}{2}}.`$ (85)
One can see that for $`d=1`$ which would correspond to the canonical dimension of a scalar field the scalar mass falls much faster, for a $`\mathrm{\Phi }(\rho )`$ that decreases as a function of density, than what would be given by BR scaling. Increasing the $`d`$ (and hence the anomalous dimension) makes the scalar mass fall less rapidly. With $`d2`$, $`\kappa _d1`$ and we recover the BR scaling. Since the dropping scalar mass is associated with an increasing attraction, we see that the anomalous dimension plays the role of bringing in an effective repulsion. One may therefore interpret this as a screening effect of the scalar attraction. In particular, that $`d2.7>0`$ means that in FTS1, an additional screening of the BR-scaled scalar exchange (or an effective repulsion) is present.
## 4 Brown-Rho scaling
Brown and Rho develop a strategy of density-dependent effective field theory for an in-medium effective theory in . They assume that the effective Lagrangian even for the hadrons in matter also keeps the symmetries of QCD (e.g., chiral symmetry) and that the parameters of the effective theory are determined at a given density. The change of vacuum in the presence of medium is assumed to be expressible by the change of parameters of the theory. Brown-Rho scaling is the relation among the density-dependent parameters in medium. The quasiparticle picture associated with BR scaling is a successful way to describe hadrons in medium. In this section BR scaling is derived from QCD-oriented effective Lagrangian and discussed.
### 4.1 Freund-Nambu model
Before discussing the effective theory for dense matter and BR scaling, we study in this section the Freund-Nambu (FN) model, in order to describe the idea of BR scaling. The FN model is a simple model where scale symmetry is realized in Goldstone mode. The FN model Lagrangian is
$`^{FN}`$ $`=`$ $`_{inv}^{FN}+_{sb}^{FN}`$ (86)
$`_{inv}^{FN}`$ $`=`$ $`{\displaystyle \frac{1}{2}}^\mu \varphi _\mu \varphi +{\displaystyle \frac{1}{2}}^\mu \chi _\mu \chi {\displaystyle \frac{1}{2}}f^2\chi ^2\varphi ^2`$
$`_{sb}^{FN}`$ $`=`$ $`{\displaystyle \frac{c}{8}}(\chi ^2v^2)^2`$
with matter field $`\varphi `$ and dilaton field $`\chi `$ both of which have scale dimension 1;
$`\delta \varphi =ฯต(1+x)\varphi \delta \chi =ฯต(1+x)\chi .`$ (87)
Equation (87) comes from $`\mathrm{\Phi }(ax)=a^1\mathrm{\Phi }(x)`$ with $`a=1+ฯต`$ where $`\mathrm{\Phi }`$ represents fields of scale dimension 1. The equations of motion for the FN model
$`^2\varphi f^2\chi ^2\varphi `$ $`=`$ $`0`$ (88)
$`^2\chi f^2\chi \varphi ^2{\displaystyle \frac{c}{2}}\chi (\chi ^2v^2)`$ $`=`$ $`0`$ (89)
have non-trivial constant solutions;
$`\varphi =0,\chi =\pm v.`$ (90)
Let $`\chi =v`$ and shift the dilation field as
$`\chi =\chi ^{}+v.`$ (91)
The scale transformation of the newly defined $`\chi ^{}`$ does not show a definite scale dimension, but show a symptom of Goldstone mode;
$`\delta \chi ^{}=ฯต(1+x)\chi ^{}+ฯตv.`$ (92)
With the field redefinition
$`\sigma {\displaystyle \frac{v}{2}}({\displaystyle \frac{\chi ^2}{v^2}}1)`$ (93)
Eq. (86) becomes
$`_{inv}^{FN}`$ $`=`$ $`{\displaystyle \frac{1}{2}}^\mu \varphi _\mu \varphi {\displaystyle \frac{m_\varphi ^2}{2}}\varphi ^2{\displaystyle \frac{m_\varphi ^2}{v}}\sigma \varphi ^2+{\displaystyle \frac{1}{2}}^\mu \sigma _\mu \sigma {\displaystyle \frac{1}{1+2\chi /v}}`$
$`_{sb}^{FN}`$ $`=`$ $`{\displaystyle \frac{m_\sigma ^2}{2}}\sigma ^2`$ (94)
with
$`m_\varphi `$ $`=`$ $`f\chi `$ (95)
$`m_\sigma `$ $`=`$ $`\sqrt{c}\chi .`$
Scale symmetry in $`_{inv}`$ becomes invisible.
We can see some interesting features in the FN model. Both the matter and dilaton field masses depend on $`\chi `$. It means that the both masses move in the same way if the vacuum expectation value of $`\chi `$ is changed. In addition, the $`\sigma `$ mass can be left massive even if $`c0`$. We must also note that the explicit scale symmetry breaking term is necessary for the FN model picture. It makes the spontaneously scale symmetry breaking of $`_{inv}`$ possible.
### 4.2 Brown-Rho scaling in in-medium effective Lagrangian
We can introduce the elementary scalars that QCD has. As Schechter did , the scalar fields represented by the trace anomaly (60) can be introduced. We divide the trace anomaly into a hard part, which is associated with the gluonium $`\chi _g`$, and a soft part, which comes from the quarkonium $`\chi `$.
$`\theta _\mu ^\mu =(\theta _\mu ^\mu )_{hard}+(\theta _\mu ^\mu )_{soft}.`$ (96)
The gluon condensate in $`(\theta _\mu ^\mu )_{hard}`$ determines the gluonium mass $`m_{\chi _g}1.61.8`$ GeV and does not vanish . Since gluonium mass is much larger than the chiral scale $`\mathrm{\Lambda }_\chi 1`$ GeV, gluonium fields are integrated out in low energy physics. And Adami and Brown show by QCD sum rule that the gluon condensate is less important for the masses of light-quark hadrons. So we focus on the quarkonium scalar which has mostly to do with $`(\theta _\mu ^\mu )_{soft}`$.
In matter-free space, there seems to be no scalar whose mass is small enough, i.e.,$`\mathrm{\Lambda }_\chi `$. However, according to Weinbergโs mended symmetry there must be a scalar to form a quartet with pions near the chiral phase transition. In addition, lattice simulations near the chiral phase transition shows that two light flavor QCD transition reproduces a scaling relation with O(4) exponents as argued first by Pisarski and Wilczek . Beane and van Kolck suggest that the Goldstone boson from spontaneous scale symmetry breaking plays the role of the chiral partner of the pion, i.e., the chiral singlet scalar in the scale anomaly approaches the pions and makes up the quartet of O(4) symmetry in medium as density increases to the critical density of the chiral restoration. And we must note that the quark condensate, which contributes to the quarkonium mass, goes to zero as chiral symmetry is restored. So we assume that the quarkonim mass in the scale anomaly becomes $`m_\chi ^{}\mathrm{\Lambda }_\chi `$ as density increases, though $`m_\chi \mathrm{\Lambda }_\chi `$ in free space.
Integrating out the hard part of the scalar field $`\chi _g`$, we introduce into a Skyrme Lagrangian - which represents the low energy QCD for the infinite number of colors - the soft part of scalar, $`\chi `$, in order to make our effective theory consistent with QCD scale property.
$``$ $`=`$ $`_{inv}+_{SB}`$ (97)
$`_{inv}`$ $`=`$ $`{\displaystyle \frac{f_\pi ^2}{4}}\left({\displaystyle \frac{\chi }{\chi _0}}\right)^2\mathrm{Tr}(_\mu U^\mu U^{})+{\displaystyle \frac{1}{2}}_\mu \chi ^\mu \chi `$ (98)
$`+{\displaystyle \frac{1}{32e^2}}\mathrm{Tr}[U^{}_\mu U,U^{}_\nu U]^2+\mathrm{}`$
$`_{SB}`$ $`=`$ $`V(\chi )+\text{pion mass term}+\mathrm{}`$ (99)
with
$`U`$ $`=`$ $`e^{\frac{i๐๐
}{f_\pi }}`$ (100)
$`\chi _0`$ $`=`$ $`0|\chi |0.`$ (101)
We make $`_{inv}`$ scale correctly by multiplying the proper powers of $`\chi `$. Since $`_{inv}`$ do not contribute to $`\theta _\mu ^\mu `$, we add the scale breaking potential term $`V(\chi )`$ due to scale anomaly. $`V(\chi )`$ includes radiative corrections of high order and gives the trace anomaly of QCD in terms of $`\chi `$.
Let us break the scale symmetry spontaneously following the strategy of the FN model. If we define $`\chi _{}`$ as the mean field value of $`\chi `$ in dense matter
$`\chi _{}\chi _\rho ,`$ (102)
we can expand $`\chi `$ as
$`\chi =\chi _{}+\chi ^{}.`$ (103)
Stars will be affixed to all the quantities that appear in nuclear matter from now on. $`_{inv}`$ becomes in terms of $`\chi ^{}`$,
$`_{inv}`$ $`=`$ $`{\displaystyle \frac{f_\pi ^2}{4}}\mathrm{Tr}(_\mu U^\mu U^{})+{\displaystyle \frac{1}{2}}_\mu \chi ^{}^\mu \chi ^{}`$ (104)
$`+{\displaystyle \frac{1}{32e^2}}\mathrm{Tr}[U^{}_\mu U,U^{}_\nu U]^2+\mathrm{}`$
with the effective pion decay constant defined at mean field level as
$`f_\pi ^{}=f_\pi {\displaystyle \frac{\chi _{}}{\chi _0}}.`$ (105)
Note that $`U^{}=e^{i๐๐
^{}/f_\pi ^{}}`$ with $`๐
^{}\pi \chi _{}/\chi _0`$ is the same as $`U`$ in (100) by the definition (105). In our effective field theory both scale symmetry and chiral symmetry are realized in the Goldstone mode and their Goldstone bosons are $`\chi `$ and $`\pi `$โs, respectively.
According to Gell-Mann-Oaks-Renner (GMOR) relation
$`f_\pi ^2m_\pi ^2={\displaystyle \frac{m_u+m_d}{2}}0|\overline{u}u+\overline{d}d|0,`$ (106)
The pion mass is proportional to the quark mass. Since the quark masses comes from explicit chiral symmetry breaking which has something to do with the electroweak symmetry breaking scale $`\mathrm{\Lambda }_{EW}\mathrm{\Lambda }_\chi `$, the pion mass problem is out of our interesting range. In the prediction of chiral perturbation theory in medium the pole mass of the pion does not decrease up to nuclear matter density. In fact a recent analysis of deeply bound pionic states in heavy nuclei shows that the pole mass of the pion could be even a few per cents higher than the free space value at nuclear matter density. The $`m_\pi ^{}`$ in our in-medium effective chiral Lagrangian is not necessarily the pole mass. Thus it is not clear how to incorporate this empirical information into our theory. We will assume here that $`m_\pi ^{}`$ does not scale. This assumption may be justified by using the fact that the pion is an almost perfect Goldstone boson. Under the assumption $`m_\pi ^{}=m_\pi `$, Eq. (106) may give
$`{\displaystyle \frac{f_\pi ^{}}{f_\pi }}\left({\displaystyle \frac{\overline{q}q^{}}{\overline{q}q}}\right)^{1/2}.`$ (107)
It is an example to show the relation between the BR scaling factor and the change of the vacuum in medium.
Since the hedgehog solution of the Skyrme Lagrangian (98) gives the Skyrmion mass proportional to $`\sqrt{g_A}f_\pi `$, the nucleon mass must scale in the matter as
$`{\displaystyle \frac{m_N^{}}{M}}=\sqrt{{\displaystyle \frac{g_A^{}}{g_A}}}{\displaystyle \frac{f_\pi ^{}}{f_\pi }}.`$ (108)
Here $`M`$ represents the nucleon mass in free space and $`m_N^{}`$ is the in-medium effective mass of nucleon (later identified with the Landau effective mass). In the simple Skyrme model $`g_A`$ is inversely proportional to $`e^2`$ which does not scale since the quartic Skyrme term is classically scale-invariant. So $`g_A`$ does not change in our simple model and the nucleon mass scales in the same way as pion decay constant in medium.
The successful low energy results of the tree level in the chiral effective theory implemented with hidden local symmetry, i.e., Kawarabayashi-Suzuki-Riazuddin-Fayyazudin(KSRF) relation, $`\rho `$ coupling universality, $`\rho `$-meson dominance, etc., are shown to remain valid with the loop effects at low energy . Though it is proven in free space, it might hold in medium. If we assume KSRF relation
$`m_v^2=2e^2f_\pi ^2`$ (109)
is satisfied in nuclear matter, we obtain
$`{\displaystyle \frac{m_v^{}}{m_v}}={\displaystyle \frac{f_\pi ^{}}{f_\pi }}`$ (110)
with the subscript $`v`$ standing for light-quark vector mesons $`\rho `$ and $`\omega `$.
The fluctuating component $`\chi ^{}`$ in the soft $`\chi `$ is expected to represent multi-pion excitations in scalar channel and to give a scalar effective field $`\sigma `$ in dense medium by interpolation. It is known that the correlated 2$`\pi `$ exchange can be approximated by a scalar field with a broad mass distribution . Durso, Kim, and Wambachโs recent calculation of $`N\overline{N}\pi \pi `$ helicity amplitude in the scalar channel $`f_+^{J=0}`$ for $`\rho \rho _0`$ with BR scaling of vector meson mass shows that the resonance of the scalar mass becomes very sharper and that its value shifts downward $``$ 500 MeV . $`\rho _0`$ represents the normal nuclear matter density (0.16/fm<sup>3</sup>). The light ($`\mathrm{\Lambda }_\chi `$) and decreasing mass of the scalar field suggests that it is the expected Beane and van Kolckโs dilaton . Since the main $`\pi \pi `$ rescattering comes from $`\rho `$-meson exchange , the shift of the scalar mass in medium is affected by the density-dependence of the $`\rho `$ mass. It is clear for low densities, where the linear approximation works well, that
$`{\displaystyle \frac{m_\sigma ^{}}{m_\sigma }}{\displaystyle \frac{m_v^{}}{m_v}}.`$ (111)
Now we find that the hadron masses and pion decay constant decrease in the similar way.
$`\mathrm{\Phi }(\rho ){\displaystyle \frac{f_\pi ^{}(\rho )}{f_\pi }}{\displaystyle \frac{m_\sigma ^{}(\rho )}{m_\sigma }}{\displaystyle \frac{m_v^{}(\rho )}{m_v}}{\displaystyle \frac{m^{}(\rho )}{M}}.`$ (112)
For the moment, we are ignoring the scaling of $`g_A`$ to which we will return later. $`M^{}`$ represents the effective nucleon mass obtained with a fixed $`g_A^{}`$. The universal factor $`\mathrm{\Phi }(\rho )`$ can be determined by experiments and/or QCD sum rules.
Note that the scalar field that governs BR scaling is the quarkonium component of the scale anomaly, not the hard gluonic component. The latter gives dominant contribution to the scale anomaly in QCD but is integrated out in the effective Lagrangian. This structure imposes the hadron scaling relation (112) by a Nambu-Jona-Lasinio(NJL) mechanism as described in . Recently Liu et al.โs detailed lattice analysis for the source of the mass of a constituent quark supports this structure. They show that dynamical symmetry breaking contribution gives most of the mass of the chiral constituent quark. It means that the change of the vacuum in (112) can be related only in a subtle way to the light quark hadron mass.
When considering BR scaling, one must note that the BR-scaled masses and decay constants are the parameters in an effective theory. An effective parameter defined in one theory can be different from the parameter defined in another theory, even though two theories have the same physics in common. So the connection between the BR-scaled parameters and the parameters in the other theories needs be established via observed quantities, i.e., experiments.
One must also note that BR scaling is the result of mean field approximation. So when one considers the cases where higher order modifications are important, that is, the mean-field approximation is not reliable, BR scaling cannot be applied without further corrections. Let us take Goldberger-Treiman relation
$`g_{\pi NN}f_\pi =g_AM`$ (113)
as an example. In the real world $`g_A`$ decreases from $``$ 1.26 in free space to $``$ 1 at normal nuclear matter density since it is affected by the short range interaction between baryons. So the naive BR scaling, which would have $`g_{\pi NN}`$ remain constant, can not explain Goldberger-Treiman relation in the low density region. It means higher order modifications spoil the tree-order BR scaling at low densities. At high densities above the normal nuclear matter density, however, $`g_A^{}`$ remains at 1 while $`f_\pi ^{}`$ continues to drop and hence the coupling constant ratio increases.
Although Brown and Rhoโs arguments about the scaling relation of effective parameters may seem a bit drastic, the experiments (e.g., CERES , HELIOS-3 ) provide support for this scheme. The explanation of CERES data is one of the good examples. As seen in Fig. 7, the scaling of effective masses of hadrons in medium reproduces the CERES results very simply at the mean field level .<sup>5</sup><sup>5</sup>5In Ref. BR scaling is realized on the basis of the constituent quark model . Li, Ko, and Brown used Walecka theory to implement the density dependence of the constituent quarks.
### 4.3 Duality
Brown-Rho scaling describes the low-mass dilepton enhancement on the basis of the quasiquark picture. But the increase of dilepton yields can also be described by a hadronic theory with free masses, introducing the appropriate variables, e.g. nucleonic excitations. Rapp, Chanfray and Wambach described the CERES results successfully using conventional many-body theory. To evaluate the in-medium rho meson propagator they renormalized the intermediate two-pion states, which dress rho mesons and interact with the surrounding nucleons and deltas, and considered the direct interaction of rho mesons with surrounding hadrons, especially $`\rho `$-like baryon-hole excitations (โrhosobarโ). They found that such medium effects broaden the spectral function of rho meson and the dilepton production at $`m_\rho /2`$ is enhanced.
Rapp et al.โs hadronic rescattering approach and Brown-Rhoโs quasiparticle approach have merits and demerits, compared with each other. Since BR scaling is approximate at mean-field level, Rapp et al.โs theory is more reliable at low densities than BR scaling. On the other hand at higher densities many diagrams have to be considered in the hadronic rescattering approach. BR scaling gives a medium-modified vacuum around which the weak fluctuations can be dealt with.
If both descriptions are correct, the effective variables are expected to change smoothly from hadrons to quasiquarks subject to BR scaling and both descriptions must show duality around the hadron-quark changeover densities<sup>6</sup><sup>6</sup>6Rapp and Wambach suggested /citerw99 recently interpreting the strong broadening of the $`\rho `$-meson resonance as a reminiscence of hadron-quark changeover. Such duality was suggested by Brown et al and Y. Kim et al studied it more precisely. For rho mesons in medium they studied a two-level model, which consists of the collective rhosobar \[$`N^{}(1520)N^1`$\]<sup>7</sup><sup>7</sup>7$`N^{}(1520)`$ gives the most important contribution to the photoabsorption cross sections in the dileption analysis . and the โelementaryโ $`\rho `$.
The in-medium $`\rho `$-meson propagator is
$`D_\rho =\left[q_0^2๐^2(m_\rho ^0)^2\mathrm{\Sigma }_{\rho \pi \pi }\mathrm{\Sigma }_{\rho BN}\right]^1`$ (114)
where the $`\mathrm{\Sigma }`$ indicates self-energies and $`m_\rho ^0`$ is the bare mass of $`\rho `$. Taking the free rho meson mass $`m_\rho =(m_\rho ^0)^2+\mathrm{Re}\mathrm{\Sigma }_{\rho \pi \pi }`$, we obtain the dispersion relation in the $`๐=0`$ limit
$`q_0^2=m_\rho ^2+\mathrm{Re}\mathrm{\Sigma }_{\rho N^{}N}.`$ (115)
The phenomeological Lagrangian for the s-wave interaction<sup>8</sup><sup>8</sup>8For p-wave coupling, the Lagrangian is $`={\displaystyle \frac{f_{\rho BN}}{m_\rho }}B^{}(๐\times ๐)๐_at_aN+\text{h.c.}`$ between the elmentary $`\rho `$ and the $`\rho `$-like baryon-hole excitation is
$`_{\rho BN}={\displaystyle \frac{f_{\rho BN}}{m_\rho }}B^{}(q_0๐๐_a\rho _a^0๐๐)t_aN+\text{h.c.}`$ (116)
with appropriate spin operator $`๐`$ and isospin operators $`t_a`$ . The self-energy from the interaction (116) for $`B=N^{}(1520)`$ is
$`\mathrm{\Sigma }_{\rho N^{}(1520)N}(q_0){\displaystyle \frac{8}{3}}f_{\rho N^{}N}^2{\displaystyle \frac{q_0^2}{m_\rho ^2}}{\displaystyle \frac{\rho _n}{4}}\left({\displaystyle \frac{2\mathrm{\Delta }E}{(q_0+i\mathrm{\Gamma }_t/2)^2(\mathrm{\Delta }E)^2}}\right)`$ (117)
where $`\rho _n`$ is the nuclear density and the total width $`\mathrm{\Gamma }_t`$ includes the the free width of $`M_N^{}(1520)`$ and its modification in medium. Neglecting nuclear Fermi motion ($`๐=0`$ limit), $`\mathrm{\Delta }E=M_N^{}M_N`$.
Kim et al. showed that the dispersion relation (115) gives the two states, which have $`\rho `$-meson quantum numbers, with the spectral weight
$`Z=\left(1{\displaystyle \frac{}{q_0}}\mathrm{Re}\mathrm{\Sigma }_{\rho N^{}N}\right)^1`$ (118)
and that one of them can be interpreted as an in-medium vector meson whose mass decreases. Figure 8 shows the results with $`\mathrm{\Gamma }_t=0`$. R/W indicates that $`m_\rho `$ in (117) is the free mass and B/R indicates that $`q_0/m_\rho `$ in (117) is replaced by 1, i.e., replacing $`m_\rho `$ by $`m_\rho ^{}`$. Note that $`q_0`$ i.e., the in-medium $`\rho `$-meson mass, cannot go to zero at any density as seen in R/W of Fig. 8 if $`m_\rho `$ in the denominator of (117) is fixed. It matches neither BR scaling nor the prediction that $`m_\rho ^{}0`$ at the chiral transition point . Brown et al suggested the replacement of $`m_\rho `$ in (117) by $`m_\rho ^{}`$ in order to go from Rapp et al.โs hadronic theory to BR scaling which predicts zero vector meson mass at some high density. B/R of Fig. 8 shows that in-medium $`\rho `$-meson mass goes to zero near $`\rho 2.75\rho _0`$ as predicted in . The study of the schematic model provides how BR scaling can be interpolated from a hadronic rescattering description.
## 5 Effective Lagrangian with BR scaling
If the large anomalous dimension of the scalar field in FTS1 is a symptom of a strong-coupling regime, it suggests that one should redefine the vacuum in such a way that the fluctuations around the new vacuum become weak-coupling. This is the basis of the BR scaling introduced in . The basic idea is to fluctuate around the โvacuumโ defined at $`\rho \rho _0`$ characterized by the quark condensate $`\overline{q}q_\rho \overline{q}q^{}`$. This theory was developed with a chiral Lagrangian implemented with the trace anomaly of QCD as seen in the last section. The Lagrangian used was the one valid in the large $`N_c`$-limit of QCD and hence given entirely in terms of boson fields from which baryons arise as solitons (skyrmions): Baryon properties are therefore dictated by the structure of the bosonic Lagrangian, thereby leading to a sort of universal scaling between mesons and baryons. One can see, using a dilated chiral quark model, that BR scaling is a generic feature also at high temperature in the large $`N_c`$ limit .<sup>9</sup><sup>9</sup>9If one calculate $`f_\pi ^{}(T)`$ and $`m_N^{}(T)`$ for zero density in chiral perturbation theory, the temperature-dependence deviates from BR scaling at low $`T`$ . In Y. Kim, H.K. Lee and M. Rho discussed the point that BR scaling holds at non-zero density.
In this description, one is approximating the complicated strong interaction process at a given nuclear matter density in terms of โquasiparticleโ excitations for both baryons and bosons in medium. This means that the properties of fermions and bosons in medium at $`\rho \rho _0`$ are given in terms of tree diagrams with the properties defined in terms of the masses and coupling constants universally determined by the quark condensates at that density.
The question then is: How can one marry the FTS1 Lagrangian with the BR-scaled Lagrangian? The next question is how to identify BR-scaled parameters with the Landau parameters. In this section we will provide some answers to these two questions.
### 5.1 A Hybrid BR-scaled model
As a first attempt to answer these questions, we consider the hybrid model in which the ground state is given by the mean field of the FTS1 Lagrangian $`_{FTS1}`$ and the fluctuation around the ground state is described by the tree diagrams of the BR-scaled Lagrangian $`\mathrm{\Delta }`$,
$`^{eff}=_{FTS1}+\mathrm{\Delta }.`$ (119)
This model with the canonical parameters (T1) for the ground state and a BR-scaled fluctuation Lagrangian in the non-strange flavor sector was recently used by Li, Brown, Lee, and Ko for describing simultaneously nucleon flow and dilepton production in heavy ion collisions. The nucleon flow is sensitive to the parameters of the baryon sector, in particular, the repulsive nucleon vector potential at high density whereas the dilepton production probes the parameters of the meson sector. With a suitable momentum dependence implemented to the FTS1 mean field equation of state, the nucleon flow comes out in good agreement with experiments. Furthermore the scaling of the nucleon mass in the FTS1 theory in dense medium, say, at $`\rho 3\rho _0`$, is found to be essentially the same as that given by the NJL model. Therefore we can conclude that the nucleon in FTS1 scales in the same way as BR scaling.
The dilepton production involves both baryon and meson properties, the former in the scaling of the nucleon mass and the latter in the scaling of the vector meson ($`\rho `$) mass. The equation of state correctly describing the nucleon flow and the BR-scaled vector meson mass is found to fit the dilepton data equally well, comparable to the fit obtained in using Walecka mean field. What is important in this process is the scalar mean field which governs the BR scaling and hence the production rate comes out essentially the same for FTS1 and Walecka mean fields. The delicate interplay between the attraction and the repulsion that figures importantly in the compression modulus does not play an important role in the dilepton process.
Let us see how the particles behave in the background of the FTS1 ground state given by $`_{FTS1}`$. The nucleon of course scales ร la BR as mentioned above. We can say nothing on the pion and the $`\rho `$ meson with the FTS1 theory. However there is nothing which would preclude the $`\rho `$ scaling ร la BR and the pion non-scaling within the scheme. What is encoded in the FTS1 theory is the behavior of the $`\omega `$ and the scalar $`S`$ which figure importantly in Walecka mean fields. Let us therefore focus on these two fields in medium near normal nuclear matter density.
We have already shown in subsection 3.5.3 that the mass of the scalar field $`S`$ drops less rapidly than BR scaling for $`d>2`$. One can think of this as a screening of the four-Fermi interaction in the scalar channel that arises when the scalar meson with the BR-scaled mass is integrated out. This feature and the property of the $`\omega `$ field can be seen by the toy model of the FTS1 Lagrangian (that includes terms corresponding up to three-body forces)
$`_{toyFTS1}=_{BR}+{\displaystyle \frac{m_\omega ^2}{2}}(2+\eta ){\displaystyle \frac{\varphi }{S_0}}\omega ^2{\displaystyle \frac{m_s^2\varphi ^3}{3S_0}}`$ (120)
where
$`_{BR}`$ $`=`$ $`\overline{N}(i\gamma _\mu (^\mu +ig_v\omega ^\mu )M+g_s\varphi )N`$ (121)
$`+{\displaystyle \frac{m_\omega ^2}{2}}\omega ^2(1{\displaystyle \frac{2\varphi }{S_0}}){\displaystyle \frac{m_s^2}{2}}\varphi ^2(1{\displaystyle \frac{2\varphi }{3S_0}}).`$
We have written $`_{BR}`$ such that the BR scaling is incorporated at mean field level as
$`\mathrm{\Phi }(\rho )={\displaystyle \frac{M^{}}{M}}={\displaystyle \frac{m_s^{}}{m_s}}={\displaystyle \frac{m_\omega ^{}}{m_\omega }}1{\displaystyle \frac{\varphi }{S_0}}`$ (122)
with
$`S_0=0|S|0=M/g_s.`$ (123)
Here we are ignoring the deviation of the scaling of the effective nucleon mass (denoted later as $`m_N^{}`$) from the universal scaling $`\mathrm{\Phi }(\rho )`$ . This will be incorporated in the next subsection. We can see from (120) that the FTS1 theory brings in an additional repulsive three-body force coming from a cubic scalar field term while if one takes $`\eta =2`$, the $`\omega `$ field will have a BR scaling mass in nuclear matter. Fit to experiments favors $`\eta 1/2`$ instead of $`2`$, thus indicating that the FTS1 theory requires a many-body suppression of the repulsion due to the $`\omega `$ exchange two-body force. (In the simple model with BR scaling that we will construct below, we shall use this feature by introducing a โrunningโ vector coupling $`g_v^{}`$ that drops as a function of density.) The effective $`\omega `$ mass may be written as
$`m_\omega ^2[1+\eta {\displaystyle \frac{\varphi _0}{S_0}}]m_\omega ^2.`$ (124)
For $`\eta <0`$, we have a falling $`\omega `$ mass corresponding to BR scaling (modulo, of course, the numerical value of $`\eta `$). In FTS1, there is a quartic term $`\omega ^4`$ which is attractive and hence increases the $`\omega `$ mass. In fact, because of the attractive quartic $`\omega `$ term, we have
$`{\displaystyle \frac{m_\omega ^{}}{m_\omega }}1.12`$ (125)
at the saturation density with T1 parameter set. This would seem to suggest that due to higher polynomial (many-body) effects, the $`\omega `$ mass does not follow BR scaling in medium. Furthermore the $`\omega `$ effective mass increases slowly around this equilibrium value. On the other hand Klingl, Waas, and Weiseโs recent sum rule analysis on current correlation function follows BR scaling. It shows that effective $`\omega `$ meson mass scales down by about 15$`\%`$ at normal nuclear matter density from its free value. We will discuss the shift of vector meson mass in medium in detail in Section 5.6.
### 5.2 Model with BR scaling
The above hybrid model suggests how to construct an effective Lagrangian model that implements BR scaling and contains the same physics as the FTS1 theory. The crucial point is that such a Lagrangian is to give in the mean field the chiral liquid soliton solution. This can be done by making the following replacements in (121):
$`Mg_s\varphi _0`$ $``$ $`M^{},`$
$`m_\omega ^2(1{\displaystyle \frac{2\varphi _0}{S_0}})`$ $``$ $`m_{\omega }^{}{}_{}{}^{2},`$
$`m_s^2(1{\displaystyle \frac{2\varphi _0}{S_0}})`$ $``$ $`m_{s}^{}{}_{}{}^{2}`$ (126)
and write
$`_{BR}`$ $`=`$ $`\overline{N}(i\gamma _\mu (^\mu +ig_v\omega ^\mu )M^{}+h\varphi )N`$ (127)
$`{\displaystyle \frac{1}{4}}F_{\mu \nu }^2+{\displaystyle \frac{1}{2}}(_\mu \varphi )^2+{\displaystyle \frac{m_{\omega }^{}{}_{}{}^{2}}{2}}\omega ^2{\displaystyle \frac{m_{s}^{}{}_{}{}^{2}}{2}}\varphi ^2`$
with
$`{\displaystyle \frac{M^{}}{M}}={\displaystyle \frac{m_\omega ^{}}{m_\omega }}={\displaystyle \frac{m_s^{}}{m_s}}=\mathrm{\Phi }(\rho ).`$ (128)
The additional term $`\overline{N}h\varphi N`$ is put in to account for the difference between the Landau mass $`m_L^{}`$ to be given later and the BR scaling mass $`M^{}`$. In the chiral Lagrangian approach with BR scaling, the difference comes through the Fock term involving non-local pion exchange. This will be discussed further in the next chapter. For simplicity we will take the scaling in the form
$`\mathrm{\Phi }(\rho )={\displaystyle \frac{1}{1+y\rho /\rho _0}}`$ (129)
with $`y=0.28`$ so as to give $`\mathrm{\Phi }(\rho _0)=0.78`$ (corresponding to $`k_F=260`$ MeV) found in QCD sum rule calculations , as well as from the in-medium GMOR relation .
Note that the Lagrangian (127) treated at the mean field level would give a Walecka-type model with the meson masses replaced by the BR scaling mass.
Now in order to describe nuclear matter in the spirit of the FTS1 theory, we introduce terms cubic and higher in $`\omega `$ and $`\varphi `$ fields to be treated as perturbations around the BR background as
$`_{nbody}=a\varphi \omega ^2+b\varphi ^3+c\omega ^4+d\varphi ^4+e\varphi ^2\omega ^2+\mathrm{}`$ (130)
where $`a`$$`e`$ are โnaturalโ (possibly density-dependent) constants to be determined. By inserting for the $`\varphi `$ and $`\omega `$ fields the solutions of the static mean field equations given by $`_{BR}`$,
$`m_{s}^{}{}_{}{}^{2}\overline{\varphi }`$ $`=`$ $`h{\displaystyle \underset{i}{}}\overline{N}_iN_i`$ (131)
$`m_{\omega }^{}{}_{}{}^{2}\overline{\omega }_0`$ $`=`$ $`g_v{\displaystyle \underset{i}{}}N_i^{}N_i`$ (132)
we see that at the mean-field level, $`_{nbody}`$ generates three- and higher-body forces with the exchanged masses density-dependent ร la BR. Note that at this point, the scaling factor $`\mathrm{\Phi }`$ and the mean field value (131) are not necessarily locked to each other by self-consistency.
As the first trial, we will consider the drastically simplified model by dropping the n-body term (130) and minimally modifying the BR Lagrangian (127). We shall do this by letting as mentioned above the vector coupling โrunโ as a function of density. For this, we use the observation made in that the nucleon flow probing higher density requires that $`g_v^{}/m_\omega ^{}`$ be independent of density at low densities and decrease slightly at high densities. We shall therefore take, to simulate this particular many-body correlation effect, the vector coupling to scale as
$`{\displaystyle \frac{g_v^{}}{g_v}}={\displaystyle \frac{1}{1+z\rho /\rho _0}}`$ (133)
with $`z`$ equal to or slightly greater than $`y`$. The $`h`$ is assumed not to scale although it is easy to take into account the density dependence if necessary.
The scaling (133) seems at odds with the prediction made with the Skyrme model where using the Skyrme model with the quartic Skyrme term inversely proportional to the coupling $`e^2`$, it was found that
$`{\displaystyle \frac{e}{e^{}}}\sqrt{{\displaystyle \frac{g_A^{}}{g_A}}}.`$
It is tempting to identify (via $`SU(6)`$ symmetry) $`e`$ with $`g_v`$ that we are discussing here since the Skyrme quartic term can formally be obtained from a hidden gauge-symmetric Lagrangian by integrating out the $`\rho `$ meson field. If this were correct, one would predict that the vector coupling increases โ and not decreases โ as density increases since we know that $`g_A^{}`$ is quenched in dense matter. This identification could be too naive and incomplete in two respects, however. First of all, this skyrmion formula is a large-$`N_c`$ relation and secondly the Skyrme quartic term embodies all short-distance physics in one dimension-four term in a derivative expansion. Thus the constant $`1/e`$ must represent a lot more than just the vector-meson ($`\rho `$) degree of freedom. Furthermore we are concerned with the $`\omega `$ degree of freedom which in a naive derivative expansion would give a six-derivative term. The BR-scaled model we are constructing should involve not only short-distance physics presumably represented by the $`1/e`$ term (consistent with the understanding that the quenching of $`g_A`$ is a short-distance phenomena) but also longer-range correlations. Therefore the qualitative difference should surprise no one.
The truncated Lagrangian that we shall consider then is
$`_{BR}`$ $`=`$ $`\overline{N}(i\gamma _\mu (^\mu +ig_v^{}\omega ^\mu )M^{}+h\varphi )N`$ (134)
$`{\displaystyle \frac{1}{4}}F_{\mu \nu }^2+{\displaystyle \frac{1}{2}}(_\mu \varphi )^2+{\displaystyle \frac{m_{\omega }^{}{}_{}{}^{2}}{2}}\omega ^2{\displaystyle \frac{m_{s}^{}{}_{}{}^{2}}{2}}\varphi ^2.`$
As suggested in , chiral in-medium Lagrangians can be brought to a form equivalent to a Walecka-type model. The scalar field appearing here transforms as a singlet, not as the fourth component of $`O(4)`$ of the linear sigma model. As it stands, the Lagrangian (134) does not look chirally invariant. This is because we have dropped pion fields which play no role in the ground state of nuclear matter. In considering fluctuations around the ground state, they (and other pseudo-Goldstone fields such as kaons) should be reinstated. The chiral singlet $`\omega `$ field and $`\varphi `$ field can be considered as auxiliary fields brought in from a Lagrangian consisting of multi-Fermion field operators via a Hubbard-Stratonovich transformation.
Since we treat density-dependent parameters, we must be careful in thermodynamic consistency. After showing the way to treat density-dependent parameters in the next section, we display the results of our model for nuclear matter properties. We will argue in the next subsection that the energy density from (5.3) is independent of the way how the parameters move as density increases.
### 5.3 Thermodynamic consistency in medium
In this subsection we address the issue of thermodynamic consistency of the Lagrangian (134) treated in the mean field approximation. For instance, it is not obvious that the presence of the density-dependent parameters in the Lagrangian does not spoil the self-consistency of the model, in particular, energy-momentum conservation in the medium and also certain relations of Fermi-liquid structure of the matter. The purpose of this section is to show that there is no inconsistency in doing a field theory with BR scaling masses and other parameters. This point has not been fully appreciated by workers in the field. The Euler-Lagrange equations of motion are in the same as the ones that arise from the field for the Lagrangian wherein the masses and constants are not BR scaling. While this procedure gives correct energy density, pressure and compression modulus, the energy-momentum conservation is not automatically assured. In fact, if one were to compute the pressure from the energy-density $``$, one would find that it does not give $`\frac{1}{3}<T_{ii}>`$ (where $`T_{\mu \nu }`$ is the conserved energy-momentum tensor and the bra-ket means the quantity evaluated in the mean-field approximation as defined before) unless one drops certain terms without justification. This suggests that it is incorrect to take the masses and coupling constants independent of fields in deriving, by Noether theorem, the energy-momentum tensor. So the question is: how do we treat the field dependence of the BR scaling masses and constants?
One possible solution to this problem is as follows. In Section 4.2, the density dependence of the Lagrangian arose as the vacuum expectation value of the scalar field $`\chi `$ that figures in the QCD trace anomaly. By vacuum we mean the state of baryon number zero modified from that of true vacuum by the strong influence of the baryons in the system. See later for more on this point. It corresponded to the condensate of a quarkonium component of the scalar $`\chi `$ with the gluonium component โ which lies higher than the chiral scale โ integrated out. It was assumed to scale in dense medium in a Skyrmion-type Lagrangian subject to chiral symmetry. Now in the language of a chiral Lagrangian consisting of the nucleonic matter field $`N`$ with other massive fields integrated out, this scalar condensate would be some function of the vacuum expectation value of $`\overline{N}N`$ or $`\overline{N}\gamma _0N`$ coming from multi-Fermion field operators mentioned above. How these four-Fermi and higher-Fermi field terms can lead to BR scaling in the framework of chiral perturbation theory was discussed in . We shall follow this strategy in this paper leaving other possibilities (such as dependence on the mean fields of the massive mesons) for later investigation. For this, it is convenient to define
$`\stackrel{ห}{\rho }u^\mu \overline{N}\gamma ^\mu N`$ (135)
with unit fluid 4-velocity
$`u^\mu ={\displaystyle \frac{1}{\sqrt{1๐^2}}}(1,๐)={\displaystyle \frac{1}{\sqrt{\rho ^2๐^2}}}(\rho ,๐)`$ (136)
with the baryon current density
$`๐=<\overline{N}๐ธN>`$ (137)
and the baryon number density
$`\rho =<N^{}N>={\displaystyle \underset{i}{}}n_i.`$ (138)
We will take $`n_i`$ to be given by the Fermi distribution function, $`n_i=\theta (k_F|๐_i|)`$ at $`T=0`$. We should replace $`\rho `$ in (134) by $`\stackrel{ห}{\rho }`$ for consistency of the model. The definition of $`\stackrel{ห}{\rho }`$ makes our Lagrangian Lorentz-invariant which will later turn out to be useful in deriving relativistic Landau formulas. With this, the Euler-Lagrange equation of motion for the nucleon field is
$`{\displaystyle \frac{\delta }{\delta \overline{N}}}`$ $`=`$ $`{\displaystyle \frac{}{\overline{N}}}+{\displaystyle \frac{}{\stackrel{ห}{\rho }}}{\displaystyle \frac{\stackrel{ห}{\rho }}{\overline{N}}}`$ (139)
$`=`$ $`[i\gamma ^\mu (_\mu +ig_v^{}\omega _\mu iu_\mu \stackrel{ห}{\mathrm{\Sigma }})M^{}+h\varphi ]N`$
$`=`$ $`0`$
with
$`\stackrel{ห}{\mathrm{\Sigma }}`$ $`=`$ $`{\displaystyle \frac{}{\stackrel{ห}{\rho }}}`$
$`=`$ $`m_\omega ^{}\omega ^2{\displaystyle \frac{m_\omega ^{}}{\stackrel{ห}{\rho }}}m_s^{}\varphi ^2{\displaystyle \frac{m_s^{}}{\stackrel{ห}{\rho }}}\overline{N}\omega ^\mu \gamma _\mu N{\displaystyle \frac{g_v^{}}{\stackrel{ห}{\rho }}}\overline{N}N{\displaystyle \frac{M^{}}{\stackrel{ห}{\rho }}}.`$
This additional term which may be related to what is referred to in many-body theory as โrearrangement termsโ plays a crucial role in what follows.<sup>10</sup><sup>10</sup>10For a recent discussion on rearrangement terms as well as thermodynamic consistency in the context of standard many-body theory, see . The notion of density-dependent parameters in many-body problems is of course an old one . The equations of motion for the bosonic fields are
$`(^\mu _\mu +m_s^2)\varphi `$ $`=`$ $`h\overline{N}N`$ (141)
$`_\nu F_\omega ^{\nu \mu }+m_\omega ^2\omega ^\mu `$ $`=`$ $`g_v^{}\overline{N}\gamma ^\mu N.`$ (142)
We start with the conserved canonical energy-momentum tensor constructed a la Noether from the Lagrangian (134):
$`T^{\mu \nu }`$ $`=`$ $`i\overline{N}\gamma ^\mu ^\nu N+^\mu \varphi ^\nu \varphi ^\mu \omega _\lambda ^\nu \omega ^\lambda `$ (143)
$`{\displaystyle \frac{1}{2}}[(\varphi )^2m_s^2\varphi ^2(\omega )^2+m_\omega ^2\omega ^22\stackrel{ห}{\mathrm{\Sigma }}\overline{N}u/N]g^{\mu \nu }.`$
We shall compute thermodynamic quantities from (143) using the mean field approximation which amounts to taking
$`N`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{V}}}{\displaystyle \underset{i}{}}a_i\sqrt{{\displaystyle \frac{E_{\kappa _i}+m_N^{}}{2E_{\kappa _i}}}}\left(\begin{array}{c}\chi \\ \frac{๐๐ฟ_i}{E_{\kappa _i}+m_N^{}}\chi \end{array}\right)\mathrm{exp}(i๐ฟ_i๐i(g_v^{}\omega _0u_0\mathrm{\Sigma }+E_i)t)`$ (146)
$`h\varphi `$ $`=`$ $`C_h^2<\overline{N}N>=C_h^2{\displaystyle \underset{i}{}}n_i{\displaystyle \frac{m_N^{}}{\sqrt{๐ฟ_i^2+m_N^2}}}`$ (147)
$`g_v^{}\omega `$ $`=`$ $`C_v^2(\rho ,๐)=C_v^2{\displaystyle \underset{i}{}}n_i(1,{\displaystyle \frac{๐ฟ_i}{\sqrt{๐ฟ_i^2+m_N^2}}})`$
where $`a_i`$ is the annihilation operator of the nucleon $`i`$, with
$`n_i=<a_i^{}a_i>,`$ (148)
and
$`\mathrm{\Sigma }`$ $`=`$ $`\stackrel{ห}{\mathrm{\Sigma }},`$ (149)
$`๐ฟ_i`$ $``$ $`๐_iC_v^2๐+๐\mathrm{\Sigma },`$ (150)
and
$`E_{\kappa _i}`$ $`=`$ $`\sqrt{๐ฟ_i^2+m_N^2}.`$ (151)
$`\chi `$ is the spinor and $`๐`$ is the Pauli matrix. We have defined
$`C_v(\stackrel{ห}{\rho }){\displaystyle \frac{g_v^{}(\stackrel{ห}{\rho })}{m_\omega ^{}(\stackrel{ห}{\rho })}}`$ (152)
and
$`C_h(\stackrel{ห}{\rho }){\displaystyle \frac{h}{m_s^{}(\stackrel{ห}{\rho })}}{\displaystyle \frac{1}{\stackrel{~}{C}_h(\stackrel{ห}{\rho })}}.`$ (153)
In this approximation, the energy density is
$``$ $`=`$ $`<T^{00}>`$
$`=`$ $`<i\overline{N}\gamma ^0^0N+{\displaystyle \frac{1}{2}}m_s^2\varphi ^2{\displaystyle \frac{1}{2}}m_\omega ^2\omega ^2+\stackrel{ห}{\mathrm{\Sigma }}\overline{N}u/N>`$
$`=`$ $`{\displaystyle \frac{1}{2}}C_v^2(\rho ^2+๐^2)+{\displaystyle \frac{1}{2}}\stackrel{~}{C}_h^2(m_N^{}M^{})^2+{\displaystyle \underset{l}{}}n_l\sqrt{๐ฟ_l^2+m_N^2}\mathrm{\Sigma }๐๐.`$
Note that the $`\mathrm{\Sigma }`$-dependent terms cancel out in the comoving frame ($`๐=0`$), so that the resulting energy-density is identical to what one would obtain from the Lagrangian in the mean field with the density-dependent parameters taken as field-independent quantities.
Given the energy density (5.3), the pressure can be calculated by (at $`T=0`$)
$`p`$ $`=`$ $`{\displaystyle \frac{E}{V}}=\rho ^2{\displaystyle \frac{/\rho }{\rho }}=\mu \rho `$ (155)
$`=`$ $`{\displaystyle \frac{1}{2}}C_v^2(\rho )\rho ^2\mathrm{\Sigma }_0\rho {\displaystyle \frac{1}{2}}\stackrel{~}{C}_h^2(\rho )(m_N^{}M^{}(\rho ))^2`$
$`{\displaystyle \frac{\gamma }{2\pi ^2}}\left(E_F({\displaystyle \frac{m_N^2}{8}}k_F{\displaystyle \frac{1}{12}}k_F^3){\displaystyle \frac{m_N^4}{8}}\mathrm{ln}[(k_F+E_F)/m_N^{}]\right)`$
where $`\mu `$ is the chemical potential โ the first derivative of the energy density with respect to $`\rho `$ in the comoving frame ($`๐=0`$):
$`\mu {\displaystyle \frac{}{\rho }}|_{๐=0}=C_v^2\rho +E_F\mathrm{\Sigma }_0`$ (156)
with $`E_F=\sqrt{k_F^2+m_N^2}`$ and $`\mathrm{\Sigma }_0=\stackrel{ห}{\mathrm{\Sigma }}_{๐=0}`$. To check that this is consistent, we calculate the pressure from the energy-momentum tensor (143) in the mean field at $`T=0`$:
$`p_t`$ $``$ $`{\displaystyle \frac{1}{3}}<T_{ii}>_{๐=0}`$
$`=`$ $`{\displaystyle \frac{1}{3}}i\overline{N}\gamma _i_iN{\displaystyle \frac{1}{2}}(m_\omega ^2\omega ^2m_s^2\varphi ^22\stackrel{ห}{\mathrm{\Sigma }}N^{}N)g_{ii}_{๐=0}`$
$`=`$ $`{\displaystyle \frac{1}{2}}C_v^2(\rho )\rho ^2{\displaystyle \frac{1}{2}}\stackrel{~}{C}_h^2(\rho )(m_N^{}M^{}(\rho ))^2\mathrm{\Sigma }_0\rho `$
$`{\displaystyle \frac{\gamma }{2\pi ^2}}\left(E_F({\displaystyle \frac{m_N^2}{8}}k_F{\displaystyle \frac{1}{12}}k_F^3){\displaystyle \frac{m_N^4}{8}}\mathrm{ln}[(k_F+E_F)/m_N^{}]\right).`$
This agrees with (155). Thus our equation of state conserves energy and momentum.
We showed that a simple effective chiral Lagrangian with BR scaling parameters is thermodynamically consistent, a point which is important for studying nuclear matter under extreme conditions. It is clear however that this does not require that the masses appearing in the Lagrangian scale according to BR scaling only. What is shown in this subsection is that masses and coupling constants could depend on density without getting into inconsistency with general constraints of chiral Lagrangian field theory. This point is important for applying (134) to the density regime $`\rho 3\rho _0`$ appropriate for the CERES dilepton experiments and also kaon production at GSI(Gesellschaft fรผr Schwerionenforschung) where deviation from the simple BR scaling of might occur.
### 5.4 Results
Based on the thermodynamic consistency of density-dependent effective theories proven at the mean field level, we check whether the model Lagrangian (134) can describe the infinite nuclear matter properties successfully. The characteristic properties we try to reproduce are compression modulus, $`m_N^{}`$, and binding energy at normal nuclear matter density, and the saturation density itself.
In Table 2, three sets of parameters are listed. We take the measured free-space masses for the $`\omega `$ and the nucleon and for the scalar $`\varphi `$ for which the free-space mass cannot be precisely given, we take $`m_s=700`$ MeV (consistent with what is argued in ) so that at nuclear matter density, it comes close to what enters in the FTS1. The resulting fits to the properties of nuclear matter are given in Table 3 for the parameters given in Table 2. These results are encouraging. Considering the simplicity of the model, the model โ in particular with the S2 and S3 set โ is remarkably close in nuclear matter to the full FTS1. The compression modulus comes down toward the low value that is currently favored. In fact, the somewhat higher value obtained here can be easily brought down to about 200 MeV without modifying other quantities if one admits a small admixture of the residual many-body terms (130), as we shall shortly show. The effective nucleon Landau mass $`m_N^{}/M0.67`$ is in good agreement with what was obtained in QCD sum rule calculations and also in the next section (i.e., 0.69) by mapping BR scaling to Landau-Migdal Fermi-liquid theory. We shall see below that this has strong support from low-energy nuclear properties. What is also noteworthy is that the ratio $`(g_v^{}/m_\omega ^{})^2`$ forced upon us โ though not predicted โ is independent of the density (set S1) or slightly decreasing with density (sets S2 and S3), as required in the nucleon flow data as found by Li, Brown, Lee and Ko . In FTS1 theory, it is the higher polynomial terms in $`\omega `$ and $`\varphi `$ defining the mean fields that are responsible for the reduction in $``$ needed in . In Dirac-Brueckner-Hartree-Fock theory, it is found that while $``$ takes the free-space value $`_0`$ for $`\rho \rho _0`$, it decreases to $`0.64_0`$ at $`\rho 3\rho _0`$ due to rescattering terms which in our language would correspond to the many-body correlations.
The assumption that the many-body correlation terms in (130) can be entirely subsumed in the dropping vector coupling may seem too drastic. Let us see what small residual three-body and four-body terms in (130) as many-body correlations (over and above what is included in the running vector coupling constant) can do to nuclear matter properties. For convenience we rewrite (130) by inserting dimensional factors as
$`_{nbody}`$ $`=`$ $`{\displaystyle \frac{\eta _0}{2}}m_\omega ^2{\displaystyle \frac{\varphi }{f_\pi }}\omega ^2{\displaystyle \frac{\kappa _3}{3!}}m_s^2{\displaystyle \frac{\varphi ^3}{f_\pi }}`$
$`+{\displaystyle \frac{\zeta _0}{4!}}g_v^2\omega ^4{\displaystyle \frac{\kappa _4}{4!}}m_s^2{\displaystyle \frac{\varphi ^4}{f_\pi ^2}}+{\displaystyle \frac{\eta _1}{2}}m_\omega ^2{\displaystyle \frac{\varphi ^2}{f_\pi ^2}}\omega ^2`$
and demand that the coefficients $`\eta `$, $`\zeta `$ and $`\kappa `$ so defined be natural. The results of this analysis are given in Table 4 and Fig. 9 for various values of the residual many-body terms and compared with those of the truncated model (134) with S3 parameter set. The coefficients are chosen somewhat arbitrarily to bring our points home, with no attempt made for a systematic fit. (It would be easy to fine-tune the parameters to make the model as close as one wishes to FTS1 theory.) It should be noted that while the equilibrium density or Fermi momentum $`k_{eq}`$, the effective nucleon mass $`m_N^{}`$ and the binding energy $`B`$ stay more or less unchanged, within the range of the parameters chosen, from what is given by the BR-scaled model (134) with the S3 parameters, the compression modulus $`K`$ can be substantially decreased by the residual many-body terms. Figure 9 shows that as expected, lowering of the compression modulus is accompanied by softening of the equation of state at $`\rho >\rho _0`$. While the equilibrium property other than the compression modulus is insensitive to the many-body correlation terms, the equation of state at larger density can be quite sensitive to them. This is because for the generic parameters chosen, the $`m_N^{}`$ can vanish at a given density above $`\rho _0`$ at which the approximation is expected to break down and hence the resulting value cannot be trusted. The B2 and B4 models do show this instability at $`\rho >1.5\rho _0`$. (See Fig. A.1.)
It is quite encouraging that the simple minimal model (134) with BR scaling captures so much of the physics of nuclear matter. Of course, by itself, there is no big deal in what is obtained by the truncated model: It is not a prediction. What is not trivial, however, is that once we have a Lagrangian of the form (134) which defines the mean fields, then we are able to control with some confidence the background around which we can calculate fluctuations, which was the principal objective we set at the beginning of this approach. The power of the simple Lagrangian is that we can now treat fluctuations at higher densities as one encounters in heavy-ion collisions, not just at an equilibrium point. The description of such fluctuations does not suffer from the sensitivity with which the equation of state depends at $`\rho >\rho _0`$ on the many-body correlation terms (130).
The simple Lagrangian (134) embodies the effective field theory of QCD discussed by Furnstahl, Serot, and Tang anchored on general considerations of chiral symmetry. This Lagrangian should be viewed as an effective Lagrangian that results from two successive renormalization group โdecimationsโ, one leading to a chiral liquid structure at the chiral symmetry scale and the other with respect to the Fermi surface . The advantage of (134) is that it can, on the one hand, be connected to Landau Fermi-liquid fixed point theory of nuclear matter and, on the other hand, be extrapolated to the regime of hadronic matter produced under extreme conditions as encountered in relativistic heavy ion processes. It would, for instance, allow one, starting from the ground state of nuclear matter, to treat on the same footing the dilepton processes observed in CERES experiments as explained in and kaon production at SIS(Schwerionen-Synchrotron) energy and kaon condensation in dense matter relevant to the formation of compact stars as discussed in .
### 5.5 Landau Fermi-liquid properties of the BR-scaled model
The next issue we address is the connection between the mean-field theory of the chiral Lagrangian (5.2) and Landauโs Fermi-liquid fixed point theory as formulated in Section 3.3. As far as we know, this connection is the only means available to implement chiral symmetry of QCD in dense matter based on effective field theory. For this, we shall follow closely Matsuiโs analysis of Walecka model in mean field exploiting the similarity of our model to the latter.
#### 5.5.1 Quasiparticle interactions
The quasiparticle energy $`\epsilon _i`$ and quasiparticle interaction $`f_{ij}`$ are, respectively, given by first and second derivatives with respect to $`n_i`$:
$`\epsilon _i`$ $`=`$ $`{\displaystyle \frac{}{n_i}},`$ (159)
$`f_{ij}`$ $`=`$ $`{\displaystyle \frac{\epsilon _i}{n_j}}.`$ (160)
A straightforward calculation gives
$`\epsilon _i`$ $`=`$ $`C_v^2\rho +\sqrt{๐ฟ_i^2+m_N^2}+C_v\rho ^2{\displaystyle \frac{C_v}{n_i}}C_v^2๐^2{\displaystyle \frac{C_v}{n_i}}`$ (161)
$`+\stackrel{~}{C}_h(m_N^{}M^{})^2{\displaystyle \frac{\stackrel{~}{C}_h}{n_i}}\stackrel{~}{C}_h^2(m_N^{}M^{}){\displaystyle \frac{M^{}}{n_i}}\mathrm{\Sigma }๐{\displaystyle \frac{๐}{n_i}}`$
and
$`f_{ij}`$ $`=`$ $`{\displaystyle \frac{\epsilon _i}{n_j}}|_{๐=๐=0}`$ (162)
$`=`$ $`C_v^2+4C_v\rho {\displaystyle \frac{C_v}{\rho }}+{\displaystyle \frac{m_N^{}}{E_i}}{\displaystyle \frac{m_N^{}}{n_j}}+\rho ^2[({\displaystyle \frac{C_v}{\rho }})^2+C_v{\displaystyle \frac{^2C_v}{\rho ^2}}]`$
$`+(m_N^{}M^{})^2[({\displaystyle \frac{\stackrel{~}{C}_h}{\rho }})^2+\stackrel{~}{C}_h{\displaystyle \frac{^2\stackrel{~}{C}_h}{\rho ^2}}]+2\stackrel{~}{C}_h{\displaystyle \frac{\stackrel{~}{C}_h}{\rho }}(m_N^{}M^{}){\displaystyle \frac{}{n_j}}(m_N^{}M^{})`$
$`2\stackrel{~}{C}_h{\displaystyle \frac{\stackrel{~}{C}_h}{\rho }}(m_N^{}M^{}){\displaystyle \frac{M^{}}{\rho }}\stackrel{~}{C}_h^2{\displaystyle \frac{M^{}}{\rho }}{\displaystyle \frac{}{n_j}}(m_N^{}M^{})\stackrel{~}{C}_h^2(m_N^{}M^{}){\displaystyle \frac{^2M^{}}{\rho ^2}}`$
$`(C_v^2{\displaystyle \frac{\mathrm{\Sigma }_0}{\rho }}){\displaystyle \frac{๐_i}{E_i}}{\displaystyle \frac{๐}{n_j}}`$
with $`E_i=\sqrt{๐_i^2+m_N^2}`$. Note that $`C_v`$, $`\stackrel{~}{C}_h`$, and $`M^{}`$ are functions of
$`\stackrel{ห}{\rho }=u_0\rho ๐๐`$ (163)
in the mean field approximation. In arriving at (162), we have used the observation that in the limit $`๐0`$, we have
$`{\displaystyle \frac{u_0}{n_i}}`$ $``$ $`0,`$
$`{\displaystyle \frac{^2u_0}{n_in_j}}`$ $``$ $`{\displaystyle \frac{1}{\rho ^2}}{\displaystyle \frac{๐}{n_i}}{\displaystyle \frac{๐}{n_j}},`$
$`{\displaystyle \frac{๐}{n_i}}`$ $``$ $`{\displaystyle \frac{1}{\rho }}{\displaystyle \frac{๐}{n_i}},`$
$`{\displaystyle \frac{\stackrel{ห}{\rho }}{n_i}}`$ $``$ $`1,`$
$`{\displaystyle \frac{^2\stackrel{ห}{\rho }}{n_in_j}}`$ $``$ $`{\displaystyle \frac{1}{\rho }}{\displaystyle \frac{๐}{n_i}}{\displaystyle \frac{๐}{n_j}}`$
and that if $`f`$ is taken to be a function of the expectation value of $`\stackrel{ห}{\rho }`$, then as $`๐0`$, we have
$`{\displaystyle \frac{f}{n_i}}`$ $`=`$ $`{\displaystyle \frac{f}{\stackrel{ห}{\rho }}}{\displaystyle \frac{\stackrel{ห}{\rho }}{n_i}}`$ (164)
$``$ $`{\displaystyle \frac{f}{\rho }}`$
$`{\displaystyle \frac{^2f}{n_in_j}}`$ $`=`$ $`{\displaystyle \frac{^2f}{\stackrel{ห}{\rho }^2}}{\displaystyle \frac{\stackrel{ห}{\rho }}{n_i}}{\displaystyle \frac{\stackrel{ห}{\rho }}{n_j}}+{\displaystyle \frac{f}{\stackrel{ห}{\rho }}}{\displaystyle \frac{^2\stackrel{ห}{\rho }}{n_in_j}}`$ (165)
$``$ $`{\displaystyle \frac{^2f}{\rho ^2}}{\displaystyle \frac{1}{\rho }}{\displaystyle \frac{f}{\rho }}{\displaystyle \frac{๐}{n_i}}{\displaystyle \frac{๐}{n_j}}.`$
In the absence of the baryon current, $`๐=0`$, the quantities $`\frac{m_N^{}}{n_j}`$ and $`\frac{๐}{n_j}`$ simplify to
$`{\displaystyle \frac{m_N^{}}{n_j}}={\displaystyle \frac{\frac{M^{}}{\rho }2C_h\frac{C_h}{\rho }_ln_l\frac{m_N^{}}{E_l}C_h^2\frac{m_N^{}}{E_j}}{1+C_h^2_ln_l\frac{๐_l^2}{E_l^{3/2}}}}`$ (166)
and
$`{\displaystyle \frac{๐}{n_j}}={\displaystyle \frac{\frac{๐_j}{E_j}}{1+(C_v^2\frac{\mathrm{\Sigma }_0}{\rho })_ln_l\frac{\frac{2}{3}๐_l^2+m_N^2}{E_l^{3/2}}}}.`$ (167)
Writing in the standard way
$`f_l=(2l+1){\displaystyle \frac{d\mathrm{\Omega }}{4\pi }P_l(\frac{๐_i๐_j}{k_F^2})f_{ij}(|๐_i|=|๐_j|=k_F)},`$ (168)
we see that the last term in (162) contributes to $`f_1`$ and the sum of the rest at the Fermi surface (i.e. $`|๐_j|=k_F`$) to $`f_0`$. So
$`F_0`$ $``$ $`{\displaystyle \frac{\lambda k_FE_F}{2\pi ^2}}f_0={\displaystyle \frac{3E_F}{k_F}}\rho f_0`$
$`=`$ $`{\displaystyle \frac{3E_F}{k_F}}\rho [C_v^2+4C_v\rho {\displaystyle \frac{C_v}{\rho }}+{\displaystyle \frac{m_N^{}}{E_F}}{\displaystyle \frac{m_N^{}}{n_j}}+\rho ^2\{({\displaystyle \frac{C_v}{\rho }})^2+C_v{\displaystyle \frac{^2C_v}{\rho ^2}}\}`$
$`+(m_N^{}M^{})^2\left\{({\displaystyle \frac{\stackrel{~}{C}_h}{\rho }})^2+\stackrel{~}{C}_h{\displaystyle \frac{^2\stackrel{~}{C}_h}{\rho ^2}}\right\}+2\stackrel{~}{C}_h{\displaystyle \frac{\stackrel{~}{C}_h}{\rho }}(m_N^{}M^{}){\displaystyle \frac{}{n_j}}(m_N^{}M^{})`$
$`2\stackrel{~}{C}_h{\displaystyle \frac{\stackrel{~}{C}_h}{\rho }}(m_N^{}M^{}){\displaystyle \frac{M^{}}{\rho }}\stackrel{~}{C}_h^2{\displaystyle \frac{M^{}}{\rho }}{\displaystyle \frac{}{n_j}}(m_N^{}M^{})\stackrel{~}{C}_h^2(m_N^{}M^{}){\displaystyle \frac{^2M^{}}{\rho ^2}}]`$
and
$`F_1`$ $``$ $`{\displaystyle \frac{\lambda k_FE_F}{2\pi ^2}}f_1`$ (170)
$`=`$ $`{\displaystyle \frac{3(C_v^2\frac{\mathrm{\Sigma }_0}{\rho })\rho }{E_F+(C_v^2\frac{\mathrm{\Sigma }_0}{\rho })\rho }}.`$
#### 5.5.2 Some relations for relativistic Fermi-liquid
Here we bridge the model (134) to relativistic Fermi-liquid theory. For it we will show that the thermodynamic properties of any model like (134), which has density-dependent parameters and is Walecka-type, can be described in terms of relativistic Landau parameters derived from the mean field approximation of the model in the same way as in Section 2.3.
First let us calculate the compression modulus K defined by
$`K9\rho {\displaystyle \frac{^2(๐=0)}{\rho ^2}}.`$ (171)
It comes out to be
$`K`$ $`=`$ $`{\displaystyle \frac{3k_F^2}{E_F}}+9\rho [C_v^2+4C_v\rho {\displaystyle \frac{C_v}{\rho }}+{\displaystyle \frac{m_N^{}}{E_F}}{\displaystyle \frac{m_N^{}}{\rho }}+\rho ^2\{({\displaystyle \frac{C_v}{\rho }})^2+C_v{\displaystyle \frac{^2C_v}{\rho ^2}}\}`$
$`+(m_N^{}M^{})^2\{({\displaystyle \frac{\stackrel{~}{C}_h}{\rho }})^2+\stackrel{~}{C}_h{\displaystyle \frac{^2\stackrel{~}{C}_h}{\rho ^2}}\}+2\stackrel{~}{C}_h{\displaystyle \frac{\stackrel{~}{C}_h}{\rho }}(m_N^{}M^{}){\displaystyle \frac{}{\rho }}(m_N^{}M^{})`$
$`2\stackrel{~}{C}_h{\displaystyle \frac{\stackrel{~}{C}_h}{\rho }}(m_N^{}M^{}){\displaystyle \frac{M^{}}{\rho }}\stackrel{~}{C}_h^2{\displaystyle \frac{M^{}}{\rho }}{\displaystyle \frac{}{\rho }}(m_N^{}M^{})\stackrel{~}{C}_h^2(m_N^{}M^{}){\displaystyle \frac{^2M^{}}{\rho ^2}}]`$
Comparing (5.5.1) and (5.5.2), we obtain Eq. (29);
$`K={\displaystyle \frac{3k_F^2}{E_F}}(1+F_0).`$ (173)
In our model
$`k_F\left({\displaystyle \frac{k_i}{\epsilon _i}}\right)_{k=k_F,๐=0}=E_F\sqrt{k_F^2+m_N^2}.`$ (174)
It is verified that our model satisfies the relativistic Landau Fermi-liquid formula for the compression modulus (29).
As shown in Section 2.3, the relativistic Landau liquid satisfies the mass relation (40);
$`k_F\left({\displaystyle \frac{k_i}{\epsilon _i}}\right)_{k=k_F,\stackrel{}{v}=0}=\mu (1+F_1/3).`$
One can see from Eqs. (156), (170), and (174) that (40) is satisfied exactly in our model.
And lastly the relativistic relation for the first sound velocity (2.3.3) is satisfied automatically since (29) and (40) are satisfied in our model. So all the relativistic Landau Fermi-liquid relations in Section 2.3 are satisfied with density-dependent Walecka-type models like (134).
#### 5.5.3 Discussions
The crucial question is really how to understand the scaling masses and constants as one varies temperature and density as considered in . If one takes the basic assumption that the chiral Lagrangian in the mean field with BR scaling parameters corresponds to Landauโs Fermi-liquid fixed point theory, then one should consider first fixing the Fermi momentum $`k_F`$ and let renormalization group flow come to the fixed points of the effective mass $`m_L^{}`$ for the nucleon and Landau parameters $``$ . In this case, the scaling quantities would seem to be dependent upon $`\mathrm{\Lambda }/k_F`$, not on the fields entering into the effective Lagrangian. This paper however shows that if one wants to approach the Fermi-liquid fixed point theory starting from an effective chiral Lagrangian of QCD, it is necessary to take into account the fact that the scaling arises from the effect of multi-Fermi interactions figuring in chiral Lagrangians as implied by chiral perturbation theory described in . This is probably due to the fact that we are dealing with two-stage โdecimationsโ in the present problem โ with the Fermi surface formed from a chiral Lagrangian as a nontopological soliton (i.e., โchiral liquidโ ) โ in contrast to condensed matter situations where one starts ab initio with the Fermi surface without worrying about how the Fermi surface is formed. Our result suggests that there will be a duality in describing processes manifesting the scaling behavior. In other words, the change of โvacuumโ by density exploited in could equally be represented by a certain (possibly infinite) set of interactions among hadrons โ e.g., four-Fermi and higher-Fermi terms in chiral Lagrangians โ canonically taken into account in many-body theories starting from the usual matter-free vacuum. A notable evidence may be found in the two plausible explanations of the low mass enhancement in CERES dilepton yields in terms of scaling vector meson masses and in terms of hadronic interactions giving rise to increased widths . Recently the relation between two descriptions are discussed by Brown et al and also by Kim et al. . It is argued that the description based on the reaction dynamics and on the meson spectral function should be reliable at low density where the effective degrees of freedom are hadrons and have no contradiction with the description on BR scaling. There should however be a โcrossoverโ region at higher density at which BR scaling will become more efficient or we should go over to constituent quarks. How to relate the two description in the โcrossoverโ density regime remains an open problem.
In discussing the properties of dense matter, such as the BR scaling of masses and coupling constants, e.g., $`f_\pi ^{}`$ , we have been using a Lagrangian which preserves Lorentz invariance. This seems to be at odds with the fact that the medium breaks Lorentz symmetry. One would expect for instance that the space and time components of a current would be characterized by different constants. Specifically such quantities as $`g_A`$, $`f_\pi `$ etc. would be different if they were associated with the space component or time component of the axial current. So a possible question is: How is the medium-induced symmetry breaking accommodated in the formalism which will be discussed in the next section?
Section 5.3 and this section provide the answer to this question. Here the argument is given in an exact parallel to Walecka mean field theory of nuclear matter. One writes an effective Lagrangian with all symmetries of QCD which in the mean field defines the parameters relevant to the state of matter with density. The parameters that become constants (masses, coupling constants etc.) at given density are actually functionals of chiral invariant bilinears in the nucleon fields. When the scalar field $`\varphi `$ and the bilinear $`\psi ^{}\psi `$, where $`\psi `$ is the nucleon field, develop a non-vanishing expectation value Lorentz invariance is broken and the time and space components of a nuclear current pick up different constants. This is how for instance the Gamow-Teller constant $`g_A`$ measured in the space component of the axial current is quenched in medium while the axial charge measured in the axial charge transitions is enhanced as described in the next chapter. If one were to calculate the pion decay constant in medium, one would also find that the quantity measured in the space component is different from the time component. The way Lorentz-invariant Lagrangians figure in nuclear physics is in some sense similar to what happens in condensed matter physics.
### 5.6 Mesons in medium
It should be recalled that we extracted the scaling parameter $`\mathrm{\Phi }`$ from the in-medium property of the vector mesons. Here we will present evidences for the predicted scaling in the meson masses. There are some preliminary experimental indications for decrease in matter of the $`\rho `$ meson mass in recent nuclear experiments but we expect more definitive results from future experiments at GSI and other laboratories. In fact, this is currently a hot issue in connection with the recent dilepton data coming from relativistic heavy ion experiments at CERN(European Organization for Nuclear Research).
When heavy mesons such as the vector mesons $`\rho `$, $`\omega `$ and the scalar $`\sigma `$ are reinstated in the chiral Lagrangian, then the mass parameters of those particles in the Lagrangian, when written in a chirally invariant way, are supposed to appear with star and are assumed to scale according to Eq. (112). The question is: What is the physical role of these mass parameters? If we assume that the mesons behave also like quasiparticles, that is, like weakly interacting particles with the โdropping masses,โ then physical observables will be principally dictated by the tree diagrams of those particles endowed with the scaling masses. In this case, the masses figuring in the Lagrangian could be identified in some sense as โeffectiveโ masses of the particles in the matter. This line of reasoning was used in the work of Li, Ko and Brown to interpret the low mass enhancement of the CERES data . As discussed in Section 5.2, this treatment is consistent with an effective Lagrangian which in the mean field approximation yields the nuclear matter ground state as well as fluctuations around the ground state. The parameters of the theory, as well as their density dependence are determined by the properties of the ground state. The work of this section shows that this scheme is internally consistent. However we emphasize that the scaling we have established is for the mesons that are highly off-shell and it may not be applied to mesons that are near on-shell without further corrections (e.g., widths etc.).
Suppose one probes the propagation of an $`\omega `$ meson in nuclear medium as in HADES(High Acceptance Di-Electron Spectrometer) or TJNAF(Thomas Jefferson National Accelerator Facility) experiments, say through dilepton production. The $`\omega `$โs will decay primarily outside of the nuclear medium, but let us suppose that experimental conditions are chosen so that the leptons from the $`\omega `$ decaying inside dense matter can be detected. See for discussions on this issue. The question is whether the dileptons will probe the BR-scaled mass or the quantity given by (125). The behavior of the $`\omega `$ mass would differ drastically in the two scenarios. A straightforward application of FTS1 theory would suggest that at a density $`\rho <\rho _0`$, the $`\omega `$ mass as โseenโ by the dileptons will increase slightly instead of decrease. Since in FTS1 theory, the vector coupling $`g_v`$ does not scale, this means that $`(g_v^{}/m_\omega ^{})`$ will effectively decrease. On the other hand if the vector coupling constant drops together with the mass at increasing density as in the BR scaling model, the situation could be quite different, particularly if dileptons are produced at a density $`\rho 3\rho _0`$ as in the CERES experiments: The $`\omega `$ will then be expected to BR-scale up to the phase transition. It has been recently suggested that at some high density, Lorentz symmetry can be spontaneously broken giving rise to light $`\omega `$ mesons as โalmost Goldstoneโ bosons when a small explicit Lorentz symmetry breaking term via chemical potential is introduced. By introducing the term, they assume a state which is chirally symmetric ($`\overline{q}q=0`$) but breaks Lorentz symmetry ($`q^{}q0`$). The assumed state is metastable at $`\mu <\mu _c`$ but becomes a global minimum at $`\mu >\mu _c`$. At $`\mu >\mu _c`$, $`\omega `$ would become light but not massless due to the explicit breaking. Such mesons could be a source of copious dileptons at some density higher than normal matter density. Thus measuring the $`\omega `$ mass shift could be a key test of the BR scaling idea as opposed to the FTS1-type interpretations. This interesting issue will be studied in forthcoming experiments at GSI and TJNAF. It is interesting that the dropping $`\omega `$ mass is also found in a recent QCD sum rule calculation based on current correlation functions by Klingl, Kaiser and Weise who, however, do not see the dropping of the $`\rho `$ mass because of the large broadening of $`\rho `$ peak. If we can describe the $`\omega `$ meson in medium as a quasiparticle an $`\omega `$-nuclear bound state is feasible even in light nuclei . The process like <sup>7</sup>Li(d,<sup>3</sup>He)$`{}_{\omega }{}^{}{}_{}{}^{6}`$He is expected to be seen in GSI if such a bound state exists.
## 6 Fermi-liquid theory vs. chiral Lagrangian
### 6.1 Electromagnetic current
We will here give a brief derivation of the Landau-Migdal formula for the convection current for a particle of momentum $`๐`$ sitting on top of the Fermi sea responding to a slowly varying electromagnetic (EM) field. We will then analyze it in terms of the specific degrees of freedom that contribute to the current. This will be followed by a description in terms of a chiral Lagrangian as discussed in . This procedure will provide the link between the two approaches.
#### 6.1.1 Landau-Migdal formula for the convection current
Following Landauโs original reasoning adapted by Migdal to nuclear systems, we start with the convection current given by
$`๐ฑ={\displaystyle \underset{\sigma ,\tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}(\mathbf{}_p\epsilon _p)n_p\frac{1}{2}(1+\tau _3)}`$ (175)
where the sum goes over the spin $`\sigma `$ and isospin $`\tau `$ which in spin- and isospin-saturated systems may be written as a trace over the $`\sigma `$ and $`\tau `$ operators. More precisely, this is a matrix element of the current operator corresponding to the response to the EM field of a nucleon (proton or neutron) sitting on top of the Fermi sea. The sum over spin and isospin and the momentum integral go over all occupied states up to the valence particle. What we want is a current operator and it is deduced after the calculation is completed. One can of course work directly with the operator but the result is the same. We consider a variation of the distribution function from that of an equilibrium state
$$n_p=n_p^0+\delta n_p,$$
(176)
where the superscript $`0`$ refers to equilibrium. The variation of the distribution function induces a variation of the quasiparticle energy
$$\epsilon _p=\epsilon _p^0+\delta \epsilon _p.$$
(177)
In the equilibrium state the current is zero by symmetry, so we have
$`๐ฑ`$ $`=`$ $`{\displaystyle \underset{\sigma ,\tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}\left((\mathbf{}_p\epsilon _p^0)\delta n_p+(\mathbf{}_p\delta \epsilon _p)n_p^0\right)\frac{1}{2}(1+\tau _3)},`$ (178)
$`=`$ $`{\displaystyle \underset{\sigma ,\tau }{}}{\displaystyle }{\displaystyle \frac{d^3p}{(2\pi )^3}}((\mathbf{}_p\epsilon _p^0)\delta n_p(\mathbf{}_pn_p^0)\delta \epsilon _p)){\displaystyle \frac{1}{2}}(1+\tau _3)`$
to linear order in the variation. We consider a nucleon added at the Fermi surface of a system in its ground state. Then
$`\delta n_p={\displaystyle \frac{1}{V}}\delta ^3(๐๐){\displaystyle \frac{1\pm \tau _3}{2}}`$ (179)
and
$`\mathbf{}_pn_p^0={\displaystyle \frac{๐}{k_F}}\delta (pk_F)`$ (180)
where $`๐`$ with $`|๐|=k_F`$ is the momentum of the quasiparticle. The modification of the quasiparticle energies due to the additional particle is given by
$`\delta \epsilon _p={\displaystyle \underset{\sigma ^{},\tau ^{}}{}}{\displaystyle \frac{d^3p^{}}{(2\pi )^3}f_{p,p^{}}\delta n_{p^{}\sigma ^{}\tau ^{}}}.`$ (181)
Combining (2), (178), (179) and (181) one finds that the first term of (178) gives the operator
$`๐ฑ^{(1)}={\displaystyle \frac{๐}{m_L^{}}}{\displaystyle \frac{1+\tau _3}{2}},`$ (182)
where $`๐`$ is taken to be at the Fermi surface. The second term yields
$`\delta ๐ฑ=\delta ๐ฑ_s+\delta ๐ฑ_v={\displaystyle \frac{๐}{M}}\left({\displaystyle \frac{\stackrel{~}{F}_1+\stackrel{~}{F}_1^{}\tau _3}{6}}\right),`$ (183)
where
$`\delta ๐ฑ_s`$ $`=`$ $`{\displaystyle \frac{๐}{m_L^{}}}{\displaystyle \frac{1}{2}}{\displaystyle \frac{F_1}{3}},`$ (184)
$`\delta ๐ฑ_v`$ $`=`$ $`{\displaystyle \frac{๐}{m_L^{}}}{\displaystyle \frac{\tau _3}{2}}{\displaystyle \frac{F_1^{}}{3}}={\displaystyle \frac{๐}{m_L^{}}}{\displaystyle \frac{\tau _3}{2}}{\displaystyle \frac{F_1}{3}}+{\displaystyle \frac{๐}{m_L^{}}}{\displaystyle \frac{\tau _3}{2}}{\displaystyle \frac{F_1^{}F_1}{3}}.`$ (185)
For convenience, letโs define<sup>11</sup><sup>11</sup>11The definition of $`m_L^{}`$ is in (7)
$`\stackrel{~}{F}_l(M/m_L^{})F_l`$ (186)
with analogous definitions of $`\stackrel{~}{F}_l^{}`$, etc.. It gives another representation of (41)
$`{\displaystyle \frac{M}{m_L^{}}}=1{\displaystyle \frac{\stackrel{~}{F}_1}{3}}.`$ (187)
Putting everything together we recover the well known result of Migdal
$$๐ฑ=\frac{๐}{M}g_l=\frac{๐}{M}\left(\frac{1+\tau _3}{2}+\frac{1}{6}(\stackrel{~}{F}_1^{}\stackrel{~}{F}_1)\tau _3\right),$$
(188)
where
$$g_l=\frac{1+\tau _3}{2}+\delta g_l$$
(189)
is the orbital gyromagnetic ratio and
$$\delta g_l=\frac{1}{6}(\stackrel{~}{F}_1^{}\stackrel{~}{F}_1)\tau _3.$$
(190)
Thus, the renormalization of $`g_l`$ is purely isovector. This is due to Galilean invariance, which implies a cancellation in the isoscalar channel.
We have derived Migdalโs result using standard Fermi-liquid theory arguments. This result can also be obtained by using the Ward identity, which follows from gauge invariance of the electro-magnetic interaction. This is of course physically equivalent to the above formulation. We shall now identify specific hadronic contributions to the current (188) in two ways: the Fermi-liquid theory approach and the chiral Lagrangian approach.
#### 6.1.2 Pionic contribution
In Fermi-liquid theory approach, all we need to do is to compute the Landau parameter $`F_1`$ from the pion exchange. The one-pion-exchange contribution to the quasiparticle interaction is
$`f_{๐\sigma \tau ,๐^{}\sigma ^{}\tau ^{}}^{\pi exch.}={\displaystyle \frac{1}{3}}{\displaystyle \frac{f^2}{m_\pi ^2}}{\displaystyle \frac{๐^2}{๐^2+m_\pi ^2}}\left(S_{12}(\widehat{๐})+{\displaystyle \frac{1}{2}}(3๐๐^{})\right){\displaystyle \frac{3๐๐^{}}{2}}`$ (191)
where $`๐=๐๐^{}`$ and $`f=g_{\pi NN}(m_\pi /2M)1`$. In a relativistic formulation sketched in Appendix B, we can Fierz the one-pion exchange. Done in this way, the Fierzed scalar channel is canceled by a part of the vector channel and the remaining vector channel makes a natural contribution to the pionic piece of $`F_1`$. The one-pion-exchange contribution to the Landau parameter relevant for the convection current is
$$\frac{F_1(\pi )}{3}=F_1^{}(\pi )=\frac{3f^2m_L^{}}{8\pi ^2k_F}I_1$$
(192)
where
$`I_1={\displaystyle _1^1}๐x{\displaystyle \frac{x}{1x+\frac{m_\pi ^2}{2k_F^2}}}=2+(1+{\displaystyle \frac{m_\pi ^2}{2k_F^2}})\mathrm{ln}(1+{\displaystyle \frac{4k_F^2}{m_\pi ^2}}).`$ (193)
Note that $`F_1(\pi )`$ satisfies
$`F_1(\pi )={\displaystyle \frac{3m_L^{}}{k_F}}{\displaystyle \frac{d\mathrm{\Sigma }_\pi (p)}{dp}}|_{p=k_F}`$ (194)
with one-pion-exchange Fock contribution to the self-energy $`\mathrm{\Sigma }(p)`$ and includes the higher order contribution in $`m_L^{}`$. Thus, from Eq. (190), the one-pion-exchange contribution to the gyromagnetic ratio is
$$\delta g_l^\pi =\frac{M}{k_F}\frac{f^2}{4\pi ^2}I_1\tau _3.$$
(195)
In the next subsubsection we include contributions also from other degrees of freedom.
Letโs obtain the convection current from a chiral Lagrangian and compare it with the results given above. In absence of other meson degrees of freedom, we can simply calculate Feynman diagrams given by a chiral Lagrangian defined in matter-free space. Nonperturbative effects due to the presence of heavy mesons introduce a subtlety that will be treated below.
In the leading chiral order, there is the single-particle contribution Fig. 10a which for a particle on the Fermi surface with the momentum $`๐`$ is given by
$`๐ฑ_{1body}={\displaystyle \frac{๐}{M}}{\displaystyle \frac{1+\tau _3}{2}}.`$ (196)
Note that the nucleon mass appearing in (196) is the free-space mass $`M`$ as it appears in the Lagrangian, not the effective mass $`m_L^{}`$ that enters in the Fermi-liquid approach, (182). To the next-to-leading order, we have two soft-pion terms as discussed in . We should recall a well-known caveat here discussed already in . If one were to blindly calculate the convection current coming from Fig. 10b, there would be a gauge non-invariant term that is present because the hole line is off-shell. Figure 10c contains also a gauge non-invariant term which is exactly the same as in Fig. 10b but with an opposite sign, so in the sum of the two graphs, the two cancel exactly so that only the gauge-invariant term survives. Of course we now know that the off-shell dependence is not physical and could be removed by field redefinition ab initio. To the convection current we need, only Fig. 10b contributes
$`๐ฑ_{2body}={\displaystyle \frac{๐}{k_F}}{\displaystyle \frac{f^2}{4\pi ^2}}I_1\tau _3={\displaystyle \frac{๐}{M}}{\displaystyle \frac{1}{6}}(\stackrel{~}{F}_1^{}(\pi )\stackrel{~}{F}_1(\pi ))\tau _3.`$ (197)
We should emphasize that the Landau parameters $`\stackrel{~}{F}_1`$ and $`\stackrel{~}{F}_1^{}`$ are entirely fixed by a chiral effective Lagrangian for any density.
The sum of (196) and (197) agrees precisely with the Fermi-liquid theory result (188) and (192). This formula first derived in in connection with the Landau-Migdal parameter is of course the same as the Miyazawa formula derived nearly half a century ago. Note the remarkable simplicity in the derivation starting from a chiral Lagrangian. However, we should caution that there are some non-trivial assumptions to go with the validity of the formula. As we will see shortly, we will not have this luxury of simplicity when other degrees of freedom enter.
#### 6.1.3 Vector meson contributions and BR scaling
So far we have computed only the pion contribution to $`g_l`$. In nuclear physics, more massive degrees of freedom such as the vector mesons $`\rho `$ and $`\omega `$ of mass $`700800`$ MeV and the scalar meson $`\sigma `$ of mass $`600700`$ MeV play an important role. When integrated out from the chiral Lagrangian, they give rise to effective four-Fermion interactions:
$`_4={\displaystyle \frac{C_\varphi ^2}{2}}(\overline{N}N)^2{\displaystyle \frac{C_\omega ^2}{2}}(\overline{N}\gamma _\mu N)^2{\displaystyle \frac{C_\rho ^2}{2}}(\overline{N}\gamma _\mu \tau N)^2+\mathrm{}`$ (198)
where the coefficients $`C^{}s`$ can be identified with
$`C_M^2={\displaystyle \frac{g_M^2}{m_{M}^{}{}_{}{}^{2}}}\mathrm{with}M=\varphi ,\rho ,\omega .`$ (199)
For the moment, we make no distinction as to whether one is taking into account BR scaling or not. For the Fermi-liquid approach, this is not relevant since the parameters are not calculated. However with chiral Lagrangians, we will specify the scaling which is essential. Such interaction terms are โirrelevantโ in the renormalization group flow sense but can make crucial contributions by becoming โmarginalโ in some particular kinematic situation. A detailed discussion of this point can be found in . The effective four-Fermion interactions play a key role in stabilizing the Fermi-liquid state and leads to the fixed points for the Landau parameters. (The other fixed-point quantity, i.e. the effective mass, is put in by fiat to keep the density fixed.) In the two-nucleon systems studied in , they enter into the next-to-leading order term of the potential, which is crucial in providing the cut-off independence found for cut-off masses $`>m_\pi `$.
Again it suffices to compute the Landau parameters coming from the velocity-dependent part of heavy meson exchanges in the Fermi-liquid theory approach. We treat the effective four-Fermion interaction (198) in the Hartree approximation. Then the only velocity-dependent contributions are due to the current couplings mediated by $`\omega `$ and $`\rho `$ exchanges. The corresponding contributions to the Landau parameters are
$$F_1(\omega )=C_\omega ^2\frac{2k_F^3}{\pi ^2M}$$
(200)
and
$$F_1^{}(\rho )=C_\rho ^2\frac{2k_F^3}{\pi ^2M}.$$
(201)
The derivations of relativistic $`F_1(\omega )`$ and $`F_1^{}(\rho )`$ are shown in Appendix C.
Now the calculation of the convection current and the nucleon effective mass with the interaction (198) in the Landau method goes through the same way as in the case of the pion. The net result is just Eq. (188) including the contribution of the contact interactions (200,201), i.e.,
$`\stackrel{~}{F}_1`$ $`=`$ $`\stackrel{~}{F}_1(\pi )+\stackrel{~}{F}_1(\omega ),`$ (202)
$`\stackrel{~}{F}_1^{}`$ $`=`$ $`\stackrel{~}{F}_1^{}(\pi )+\stackrel{~}{F}_1^{}(\rho ).`$ (203)
Similarly, the nucleon effective mass is determined by (41) with
$`F_1=F_1(\pi )+F_1(\omega ).`$ (204)
In chiral Lagrangina approach the most efficient way to bring in the vector mesons into the chiral Lagrangian is to implement BR scaling in the parameters of the Lagrangian. We shall take the masses of the relevant degrees of freedom to scale in the manner of BR as (128). Note again that $`M^{}`$ is a BR scaling nucleon mass which will turn out to be different from the Landau effective mass $`m_L^{}`$ . For our purpose, it is more convenient to integrate out the vector and scalar fields and employ the resulting four-Fermi interactions (198). The coupling coefficients are modified compared to Eq. (199), because the meson masses are replaced by effective ones:
$`C_M^2={\displaystyle \frac{g_M^2}{m_{M}^{}{}_{}{}^{2}}}\mathrm{with}M=\varphi ,\rho ,\omega .`$ (205)
The coupling constants may also scale but we omit their density dependence for the moment.
The first thing we need is the relation between the BR scaling factor $`\mathrm{\Phi }`$ which was proposed in to reflect the quark condensate in the presence of matter and the contribution to the Landau parameter $`F_1`$ from the isoscalar vector ($`\omega `$) meson. For this we first calculate the Landau effective mass $`m_N^{}`$ in the presence of the pion and $`\omega `$ fields
$`{\displaystyle \frac{m_L^{}}{M}}=1+{\displaystyle \frac{1}{3}}(F_1(\omega )+F_1(\pi ))=\left(1{\displaystyle \frac{1}{3}}(\stackrel{~}{F}_1(\omega )+\stackrel{~}{F}_1(\pi ))\right)^1.`$ (206)
Next we compute the nucleon self-energy using the chiral Lagrangian. Given the single quasiparticle energy
$`\epsilon _p={\displaystyle \frac{p^2}{M^{}}}+C_\omega ^2N^{}N+\mathrm{\Sigma }_\pi (p),`$ (207)
we get the effective mass as in
$`{\displaystyle \frac{m_L^{}}{M}}={\displaystyle \frac{k_F}{m_N}}\left({\displaystyle \frac{d}{dp}}\epsilon _p|_{p=k_F}\right)^1=\left(\mathrm{\Phi }^1{\displaystyle \frac{1}{3}}\stackrel{~}{F}_1(\pi )\right)^1`$ (208)
from Eqs. (128), (186), and (194). Comparing (206) and (208), we obtain the important result
$`\stackrel{~}{F}_1(\omega )=3(1\mathrm{\Phi }^1).`$ (209)
This is an intriguing relation. It shows that the BR factor, which was originally proposed as a precursor manifestation of the chiral phase transition characterized by the vanishing of the quark condensate at the critical point , is intimately related (at least up to $`\rho \rho _0`$) to the Landau parameter $`F_1`$, which describes the quasiparticle interaction in a particular channel. We believe that the BR factor can be computed by QCD sum rule methods or obtained from current algebra relations such as the GMOR relation evaluated in medium. As was shown in , Eq. (209) implies that the BR factor governs in some, perhaps, intricate way low-energy nuclear dynamics. The equivalence discussed above between the physics of the vacuum property $`\mathrm{\Phi }`$ and that of the quasiparticle interaction $`F_1`$ due to the massive vector-meson degree of freedom suggests that the โbottom-upโ approach โ going up in density with a Lagrangian whose parameters are fixed at zero density โ and the โtop-downโ approach โ extrapolating with a Lagrangian whose parameters are fixed at some high density โ can be made equivalent at some intermediate point. If this is so in the hot and dense regime probed by relativistic heavy ion collisions, then the CERES data should also be understandable in terms of hadronic interactions without making reference to QCD variables. Because of the complexity of hadronic descriptions, it will be difficult to relate the two directly but the recent alternative explanation of the CERES data in terms of โmelting of the vector mesonsโ inside nuclear matter manifested in the increased width of the mesons due to hadronic interactions may be an indication for a possible โdualโ description at low density between what is given in QCD variables (e.g., quark condensates) and what is given in hadronic variables (e.g., the Landau parameter), somewhat reminiscent of the quark-hadron duality in heavy-light-quark systems . A possible mechanism that could make the link between the two descriptions was suggested recently by Brown et al and by Kim et al.
In the presence of the BR scaling, a non-interacting nucleon in the chiral Lagrangian propagates with a mass $`M^{}`$, not the free-space mass $`M`$. Thus, the single-particle current Fig. 10a is not given by (196) but instead by
$`๐ฑ_{1body}={\displaystyle \frac{๐}{M^{}}}{\displaystyle \frac{1+\tau _3}{2}}.`$ (210)
Now the current (210) on its own does not carry conserved charge as long as $`M^{}M`$. This means that two-body currents are indispensable to restore charge conservation. Note that the situation is quite different from the case of Fermi-liquid theory. In the latter case, the quasiparticle propagates with the Landau effective mass $`m_L^{}`$ and it is the gauge invariance that restores $`m_L^{}`$ to $`M`$. In condensed matter physics, this is related to a phenomenon that the cyclotron frequency depends on the bare mass, not on Landau effective mass. It may be referred to as Kohn effect . The bare mass in Kohn effect is restored due to the quasiparticle interactions with Galilean invariance in the same way as for the convection current in Section 6.1.1 . This clearly indicates that gauge invariance is more intricate when BR scaling is implemented. Indeed if the notion of BR scaling and the associated chiral Lagrangian are to make sense, we have to recover the charge conservation from higher-order terms in the chiral Lagrangian. This constitutes a strong constraint on the theory.
Let us now calculate the contributions from the pion and heavy meson degrees of freedom. The pion contributes in the same way as before, so we can carry over the previous result of Fig. 10b,
$`\stackrel{}{J}_{2body}^\pi ={\displaystyle \frac{๐}{M}}{\displaystyle \frac{1}{6}}(\stackrel{~}{F}_1^{}(\pi )\stackrel{~}{F}_1(\pi ))\tau _3.`$ (211)
This is of the same form as (197) obtained in the absence of BR scaling. It is in fact identical to (197) if we assume that one-pion-exchange graph does not scale in medium at least up to nuclear matter density. This assumption is supported by observations in pion-induced processes in heavy nuclei. This means that the observation that the one-pion-exchange potential does not scale implies that the constant $`g_A^{}/f_\pi ^{}`$ remains unscaling at least up to normal nuclear matter density with non-scaling pion mass. In what follows, we will make this assumption implicitly.
The contributions from the vector meson degrees of freedom are a bit trickier. They are given by Fig. 11.
Both the $`\omega `$ (isoscalar) and $`\rho `$ (isovector) channels contribute through the antiparticle intermediate state as shown in Fig. 11a. The antiparticle is explicitly indicated in the figure. However in the heavy-fermion formalism, the backward-going anti-nucleon line should be shrunk to a point as Fig. 11b, leaving behind an explicit $`1/M`$ dependence folded with a factor of nuclear density signaling the $`1/M`$ correction in the chiral expansion. One can interpret Fig. 11a as saturating the corresponding counter term although this has to be yet verified by writing the full set of counter terms at the same order. These terms with Fig. 11a
$`๐ฑ_{2body}^\omega `$ $`=`$ $`{\displaystyle \frac{๐}{M}}{\displaystyle \frac{1}{6}}\stackrel{~}{F}_1(\omega ),`$ (212)
$`๐ฑ_{2body}^\rho `$ $`=`$ $`{\displaystyle \frac{๐}{M}}{\displaystyle \frac{1}{6}}\stackrel{~}{F}_1^{}(\rho )\tau _3,`$ (213)
where $`\stackrel{~}{F}_1(\omega )`$ and $`\stackrel{~}{F}_1^{}(\rho )`$ are given by Eqs. (200,201) with $`M`$ replaced by $`M^{}`$. Their relativistic forms are given in Appendix C. The total current given by the sum of (210), (211), (212) and (213) precisely agrees with the Fermi-liquid theory result (188) when we take
$`\stackrel{~}{F}_1`$ $`=`$ $`\stackrel{~}{F}_1(\omega )+\stackrel{~}{F}_1(\pi ),`$ (214)
$`\stackrel{~}{F}_1^{}`$ $`=`$ $`\stackrel{~}{F}_1^{}(\rho )+\stackrel{~}{F}_1^{}(\pi ).`$ (215)
The way in which this precise agreement comes about is nontrivial. What happens is that part of the $`\omega `$ channel restores the BR-scaled mass $`M^{}`$ back to the free-space mass $`M`$ in the isoscalar current. (It has been known since sometime that something similar happens in the standard Walecka model (without pions and BR scaling) ). Thus, the leading single-particle operator combines with the sub-leading four-Fermi interaction to restore the charge conservation as required by the Ward identity. This is essentially the โback-flow mechanismโ which is an important ingredient in Fermi-liquid theory. We describe below the standard back-flow mechanism as given in Sec. 2.1, adapted to nuclear systems with isospin degrees of freedom, and elucidate the connection to the results obtained with the chiral Lagrangian in this subsection.
The current so constructed is valid for a process occurring very near the Fermi surface corresponding to the limit $`(\omega ,๐)(0,\mathrm{๐})`$ where $`q`$ is the spatial momentum transfer and $`\omega `$ is the energy transfer. In the diagrams considered so far (Fig. 10 and 11) the order of the limiting processes does not matter. However, the particle-hole contribution, which we illustrate in Fig. 12 with the pion contribution, does depend on the order in which $`q=|๐|`$ and $`\omega `$ approach zero. Thus, in the limit $`q/\omega 0`$, the particle-hole contributions vanish whereas in the opposite case $`\omega /q0`$, they do not.
This can be seen by examining the particle-hole propagator
$`{\displaystyle \frac{n_k(1n_{k+q})}{q_0+ฯต_kฯต_{k+q}+i\delta }}{\displaystyle \frac{n_{k+q}(1n_k)}{q_0+ฯต_kฯต_{k+q}i\delta }}`$ (216)
where $`(q_0,๐)`$ is the four-momentum of the external (EM) field. This vanishes if we set $`q0`$ with $`q_0`$ non-zero but its real part is non-zero if we interchange the limiting process since for $`q_0=0`$ we have
$`{\displaystyle \frac{๐\widehat{๐}}{๐๐/M}}\delta (k_Fk).`$ (217)
Figure 12 was computed by several authors (e.g., see ) and is given in the limit $`\omega /q0`$ by
$`๐ฑ_{ph}={\displaystyle \underset{\tau ^{}}{}}๐(1){\displaystyle \frac{1+\tau _3^{}}{2}}๐(2){\displaystyle \frac{d^3p}{(2\pi )^3}\widehat{๐}\delta (k_F|๐|)f_s^\pi }`$ (218)
where $`f_s^\pi `$ is the spatial and spin part of the quasiparticle interaction which is (191) and (B: Relativistic calculation of $`F_1^\pi `$) without isospin part $`(3๐๐^{})/2`$. The isospin factor is given by the Fierz transformation:
$`{\displaystyle \underset{\tau ^{}}{}}๐(1){\displaystyle \frac{1+\tau _3^{}}{2}}๐(2)`$ $`=`$ $`{\displaystyle \underset{\tau ^{}}{}}{\displaystyle \frac{3}{4}}{\displaystyle \frac{1}{4}}๐๐^{}+{\displaystyle \frac{3}{4}}\mathrm{tr}[\tau _3^{}]{\displaystyle \frac{1}{4}}๐\mathrm{tr}[\tau _3^{}๐^{}]`$ (219)
$`=`$ $`{\displaystyle \frac{3}{2}}{\displaystyle \frac{1}{2}}\tau _3.`$
Note that the factor $`\frac{3}{2}`$ comes from $`f_\pi `$ and $`\frac{1}{2}\tau _3`$ from $`f_\pi ^{}`$. In the limit that we are concerned with (i.e. $`T=0`$ and $`\omega /q0`$), the particle-hole contribution to the current is
$`๐ฑ_{ph}`$ $`=`$ $`{\displaystyle \frac{1}{3\pi ^2}}\widehat{๐}k_F^2(f_1+f_1^{}\tau _3)`$ (220)
$`=`$ $`{\displaystyle \frac{๐}{M}}\left({\displaystyle \frac{\stackrel{~}{F}_1(\pi )+\stackrel{~}{F}_1^{}(\pi )\tau _3}{6}}\right).`$
This holds in general regardless of what is being exchanged as long as the exchanged particle has the right quantum numbers. Contributions from heavy-meson exchanges are calculated in a similar way. Adding the particle-hole contribution (220) to the Fermi-liquid result (188) we obtain the current of a dressed or localized quasiparticle
$`๐ฑ_{locQP}={\displaystyle \frac{๐}{m_L^{}}}\left({\displaystyle \frac{1+\tau _3}{2}}\right).`$ (221)
Note that $`๐ฑ_{ph}`$ precisely cancels $`\delta ๐ฑ`$, Eq. (183). The current $`๐ฑ_{locQP}`$ is the total current carried by the wave packet of a localized quasiparticle with group velocity $`๐ฏ_F=\frac{๐}{m_L^{}}`$. However, the physical situation corresponds to homogeneous (plane wave) quasiparticle excitations. The current carried by a localized quasiparticle equals that of a homogeneous quasiparticle excitation modified by the so called back-flow current . The back-flow contribution $`(๐ฑ_{locQP}๐ฑ_{LM})`$ is just the particle-hole polarization current in the $`\omega /q0`$ limit, Eq. (220).
#### 6.1.4 Phenomenological test
It is not obvious that the effective nucleon mass computed in the chiral Lagrangian approach is directly connected to a measurable quantity although quasielastic electron scattering from nuclei does probe some kind of effective nucleon mass and Walecka model describes such a process in terms of an effective mass. To the extent that the bulk of $`m_N^{}`$ is related to the condensate through BR scaling as we can see in (208), the effective mass in the chiral Lagrangian can be related to the quantity calculated in the QCD sum rule approach for in-medium hadron masses. In BR scaling, the parameter $`\mathrm{\Phi }`$ is related to the scaling of the vector meson ($`\rho `$) mass. There are several QCD sum rule calculations for the $`\rho `$ meson in-medium mass starting with . The most recent one which closely agrees with the GMOR formula in medium for the pion decay constant $`f_\pi ^{}`$ (see Eq. (112)) is the one by Jin and Leinweber :
$`\mathrm{\Phi }(\rho _0)=0.78\pm 0.08.`$ (222)
We shall take this value in what follows but one should be aware of the possibility that this value is not quite firm. A caveat to this result was recently discussed by Klingl, Kaiser, and Weise who show that the QCD sum rule can be saturated without the mass shifting downward by increasing the vector meson width in medium. For a discussion of the empirical constraints on the in-medium widths of vector mesons, see Friman .
Given this, we can compute $`m_N^{}`$ using (208) for nuclear matter density since the pionic contribution $`\stackrel{~}{F}_1(\pi )`$ is known. One finds
$`{\displaystyle \frac{m_N^{}}{M}}(\rho =\rho _0)0.70.`$ (223)
This can be tested in an indirect way by looking at certain magnetic response functions of nuclei as described below. An additional evidence comes from QCD sum rule calculations. Again there are caveats in the QCD sum rule calculation for the nucleon mass even in free-space and certainly more so in medium. Nevertheless the most recent result by Furnstahl, Jin, and Leinweber is rather close to the prediction (223):
$`\left({\displaystyle \frac{m_N^{}(\rho _0)}{M}}\right)_{QCD}=0.69_{0.07}^{+0.14}.`$ (224)
If one writes the gyromagnetic ratio $`g_l`$ as
$`g_l={\displaystyle \frac{1+\tau _3}{2}}+\delta g_l`$ (225)
then the chiral Lagrangian prediction is
$`\delta g_l={\displaystyle \frac{1}{6}}(\stackrel{~}{F}_1^{}\stackrel{~}{F}_1)\tau _3={\displaystyle \frac{4}{9}}\left[\mathrm{\Phi }^11{\displaystyle \frac{1}{2}}\stackrel{~}{F}_1(\pi )\right]\tau _3.`$ (226)
In writing the second equality we have used (192), (209) and the nonet relation $`\stackrel{~}{F}^{}(\rho )=\stackrel{~}{F}(\omega )/9`$. At nuclear matter density, we get, using (222),
$`\delta g_l(\rho _0)0.23\tau _3.`$ (227)
This agrees with the value extracted from the dipole sum rule in <sup>209</sup>Bi ,
$`\delta g_l^{proton}=0.23\pm 0.03`$ (228)
and agrees roughly with magnetic moment data in heavy nuclei. Nuclear magnetic moments are complicated due to conventional nuclear effects. To make a meaningful comparison, one would have to extract all โtrivialโ nuclear effects and this operation brings in inestimable uncertainties. It should be stressed that the gyromagnetic ratio provides a test for the scaling nucleon mass at $`\rho \rho _0`$. It also gives a check of the relation between the baryon property on the left-hand side of Eq. (209) and the meson property on the right-hand side. Instead of using (222) as an input to calculate $`\delta g_l`$, we could take the experimental value (228) to determine, using (226), the BR scaling factor $`\mathrm{\Phi }`$ at $`\rho \rho _0`$. We would of course get (222), a value which is consistent with what one obtains in the QCD sum rule calculation and also in the in-medium GMOR relation.
Recall that because of the pions which provide (perturbative) non-local interactions to the Landau interaction, the Landau mass for the nucleon scales differently from that of the vector mesons. (See (112) and (208)). This difference is manifested in the skyrmion description by the fact that the coefficient of the Skyrme quartic term must also scale. In the original discussion of the scaling based on the quark condensates using the trace anomaly , the Skyrme quartic term was scale-invariant and hence the corresponding $`g_A^{}`$ was non-scaling. So the scaling implied by (208) indicates that the scaling of $`g_A^{}`$ is associated with the pionic degrees of freedom. This is consistent with the description based on the Landau-Migdal $`g_0^{}`$ interaction between a nucleon and a $`\mathrm{\Delta }`$ resonance and also with the QCD sum rule description of Drukarev and Levin who attribute about 80% of the quenching to the $`\mathrm{\Delta }N`$ effect.
If we equate the skyrmion relation
$`{\displaystyle \frac{m_N^{}}{M}}=\sqrt{{\displaystyle \frac{g_A^{}}{g_A}}}\mathrm{\Phi }`$ (229)
to (208), we get
$`{\displaystyle \frac{g_A^{}}{g_A}}=\left(1+{\displaystyle \frac{1}{3}}F_1(\pi )\right)^2=\left(1{\displaystyle \frac{1}{3}}\stackrel{~}{F}_1(\pi )\mathrm{\Phi }\right)^2.`$ (230)
At nuclear matter density, this predicts
$`g_A^{}(\rho _0)1`$ (231)
and
$`{\displaystyle \frac{g_A^{}}{g_A}}{\displaystyle \frac{f_\pi ^{}}{f_\pi }}=\mathrm{\Phi }.`$ (232)
We will use this relation in deriving (257). It should be emphasized that this relation, being unrelated to the vacuum property, cannot hold beyond $`\rho \rho _0`$. Indeed as suggested by the scaling given in , $`g_A^{}(\rho )\alpha g_A`$ with $`\alpha `$ a constant independent of density, for $`\rho >\rho _0`$. It would be a good approximation to set $`g_A^{}`$ equal to 1 for $`\rho >\rho _0`$.
Since one expects that when chiral symmetry is restored, $`g_A`$ will approach 1, it may be thought that the evidence for $`g_A^{}1`$ in nuclei is directly connected with chiral restoration. This is not really the case. Neither in the skyrmion picture nor in QCD sum rules is the quenching of $`g_A`$ simply related to a precursor behavior of chiral restoration. This does not however mean that the quenching of $`g_A`$ carries no information on the chiral symmetry restoration. As suggested recently by Chanfray, Delorme and Ericson , if one were to compute all pion-exchange-current graphs at one-loop order that contribute to the in-medium $`g_A`$, the effect of medium-induced change in the quark condesate would be largely accounted for. In a way, this argument is akin to that for the Cheshire-Cat (or dual) phenomenon we are advocating in the description of the quark condensate in terms of quasiparticles. Another issue that has generated lots of debate in the past and yet remains confusing is the interpretation of an effective constant $`g_A^{eff}1`$ actually observed in medium and heavy nuclei. The debate has been whether the observed โquenchingโ is due to โcore polarizationsโ or โ$`\mathrm{\Delta }`$-hole effectโ (or other non-standard mechanisms). Our view is that in the presence of BR scaling, both are involved. In light nuclei in which the Gamow-Teller transition takes place in low density, the tensor force is mainly operative and the core polarization (i.e., multiparticle-multihole configurations) mediated by this tensor force is expected to do most of the quenching, while the $`\mathrm{\Delta }`$-hole effect directly proportional to density is largely suppressed. The typical diagrammatic representations for the second order core polarization is shown in Fig. 13.
In heavy nuclei, on the other hand, the tensor force is quenched due to BR scaling, rendering the core polarization mechanism ineffective while the increased density makes the $`\mathrm{\Delta }`$-hole effect dominant. Recently Park, Jung, and Min applied chiral perturbation theory to calculate $`g_A^{}`$ at normal nuclear matter density $`\rho _0`$.
The resonance-exchange graphs that contribute are shown in Figs. 14 and 15 and the Landau-Migdal $`g_0^{}`$ effect contains both. The Hartree contribution from Fig. 14, i.e. $`\mathrm{\Delta }`$-hole effect, makes $`g_A^{}`$ quenched and the Fock contribution from Fig. 15 enhances $`g_A^{}`$. The magnitude of quenching is two or three times larger than the enhancement. What is seen in nature, in our view, is the interplay between these two.
The second form of (230) shows that the quenching of $`g_A`$ in matter is quite complex, both the pionic effect and the vacuum condensate effect being confounded together. Again for the reason given above, this relation cannot be extended beyond the regime with $`\rho \rho _0`$. We have no understanding of how this formula and the $`\mathrm{\Delta }`$-hole mechanism of are related. Our effort thus far has met with no success. Understanding the connection would presumably require the short-distance physics implied by both the Landau-Migdal $`g_0^{}`$ interaction and the Skyrme quartic term (which is known to be more than just what results when the $`\rho `$ meson is integrated out of the chiral Lagrangian).
### 6.2 Axial charge transition
No one has yet derived the analogue to (188) for the axial current. Attempts using axial Ward identities in analogy to the EM case have not met with success . The difficulty has presumably to do with the role of the Goldstone bosons in nuclear matter which is not well understood. In this subsection, we analyze the expression for the axial charge operator obtained by a straightforward application of the Fermi-liquid theory arguments of Landau and Migdal and compare this expression with that obtained directly from the chiral Lagrangian using current algebra. For the vector current we found precise agreement between the two approaches.
#### 6.2.1 Applying Landau quasiparticle argument
The obvious thing to do is to simply mimic the steps used for the vector current to deduce a Landau-Migdal expression for the axial charge operator. We use both methods developed above and find that they give the same result.
In free space, the axial charge operator nonrelativistically is $`๐๐ฏ`$ where $`๐ฏ=๐/M`$ is the velocity. In the infinite momentum frame, it is the relativistic invariant helicity $`๐\widehat{๐}`$. It is thus tempting to assume that near the Fermi surface, the axial charge operator for a local quasiparticle in a wave packet moving with the group velocity $`๐ฏ_F=๐/m_L^{}`$ is simply $`๐๐ฏ_F`$. This suggests that we take the axial charge operator for a localized quasiparticle to have the form
$`A_{0}^{}{}_{locQP}{}^{i}=g_A{\displaystyle \frac{๐๐}{m_L^{}}}{\displaystyle \frac{\tau ^i}{2}}.`$ (233)
As in the vector current case, we take (233) to be the $`\omega /q0`$ limit of the axial charge operator. The next step is to compute the particle-hole contribution to Fig. 12 (with the vector current replaced by the axial current) in the $`\omega /q0`$ limit. A simple calculation gives
$`A_{0}^{}{}_{ph}{}^{i}=g_A{\displaystyle \frac{๐๐}{m_L^{}}}{\displaystyle \frac{\tau ^i}{2}}\mathrm{\Delta }^{}`$ (234)
with
$`\mathrm{\Delta }^{}={\displaystyle \frac{f^2k_Fm_L^{}}{4m_\pi ^2\pi ^2}}(I_0I_1)`$ (235)
where $`I_1`$ was defined in (193) and
$`I_0={\displaystyle _1^1}๐x{\displaystyle \frac{1}{1x+\frac{m_\pi ^2}{2k_F^2}}}=\mathrm{ln}\left(1+{\displaystyle \frac{4k_F^2}{m_\pi ^2}}\right).`$ (236)
In an exact parallel to the procedure used for the vector current, we take the difference
$`A_{0}^{}{}_{locQP}{}^{i}A_{0}^{}{}_{ph}{}^{i}`$ (237)
and identify it with the corresponding โLandau axial chargeโ(LAC):
$`A_{0}^{}{}_{LAC}{}^{i}=g_A{\displaystyle \frac{๐๐}{m_L^{}}}{\displaystyle \frac{\tau ^i}{2}}(1+\mathrm{\Delta }^{}).`$ (238)
Let us now rederive (238) with an argument analogous to that proven to be powerful for the convection current. We shall do the calculation using the pion exchange only but the argument goes through when the contact interaction (198) is included. We begin by assuming that the axial charge โ in analogy to (175) for the convection current โ takes the form,
$`A_{0}^{}{}_{}{}^{i}=g_A{\displaystyle \underset{\sigma \tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}๐(\mathbf{}_pฯต_p)n_p\frac{\tau ^i}{2}}`$ (239)
where $`n_p`$ and $`ฯต_p`$ are $`2\times 2`$ matrices with matrix elements
$`[n_p(๐,t)]_{\alpha \alpha ^{}}=n_p(๐,t)\delta _{\alpha \alpha ^{}}+๐_p(๐,t)๐_{\alpha \alpha ^{}},`$ (240)
and
$`[ฯต_p(๐,t)]_{\alpha \alpha ^{}}=ฯต_p(๐,t)\delta _{\alpha \alpha ^{}}+๐ผ_p(๐,t)๐_{\alpha \alpha ^{}}`$ (241)
with
$`๐_p(๐,t)={\displaystyle \frac{1}{2}}{\displaystyle \underset{\alpha \alpha ^{}}{}}๐_{\alpha \alpha ^{}}[n_p(๐,t)]_{\alpha ^{}\alpha }.`$ (242)
In general $`n=4`$ in the spin-isospin space. But without loss of generality, we could confine ourselves to $`n=2`$ in the spin space with the isospin operator explicited as in Eq.(239). Then upon linearizing, we obtain from (239)
$`A_{0}^{}{}_{}{}^{i}=g_A{\displaystyle \underset{\sigma \tau }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}\left(๐(\mathbf{}_pฯต_p^0)\delta n_{p\sigma \tau }๐(\mathbf{}_pn_p^0)\delta ฯต_{p\sigma \tau }\right)\frac{\tau ^i}{2}}+\mathrm{}`$ (243)
where
$`\delta n_{p\sigma \tau }={\displaystyle \frac{1}{V}}\delta ^3(๐๐){\displaystyle \frac{1+\sigma _3}{2}}{\displaystyle \frac{\tau ^i}{2}}`$ (244)
and
$`\delta ฯต_{p\sigma \tau }={\displaystyle \underset{\sigma ^{},\tau ^{}}{}}{\displaystyle \frac{d^3p^{}}{(2\pi )^3}f_{p\sigma \tau ,p^{}\sigma ^{}\tau ^{}}\delta n_{p^{}\sigma ^{}\tau ^{}}}.`$ (245)
in analogy with (181). The equation (243) is justified if the density of polarized spins is much less than the total density of particles (assumed to hold here). The first term of (243) with (244) yields the quasiparticle charge operator
$`A_{0}^{}{}_{QP}{}^{i}=g_A{\displaystyle \frac{๐๐}{m_L^{}}}{\displaystyle \frac{\tau ^i}{2}}`$ (246)
while the second term represents the polarization of the medium, due to the pion-exchange interaction (191)
$`\delta A_{0}^{}{}_{}{}^{i}=g_A{\displaystyle \frac{๐๐}{m_L^{}}}{\displaystyle \frac{\tau ^i}{2}}\mathrm{\Delta }^{}.`$ (247)
The sum of (246) and (247) agrees precisely with the Landau charge (238).
It is not difficult to take into account the full Landau-Migdal interactions (2) which includes the one-pion-exchange interaction as well as other contributions to the quasiparticle interaction. Thus, the general expression is obtained by making the replacement
$`\mathrm{\Delta }^{}{\displaystyle \frac{1}{3}}G_1^{}{\displaystyle \frac{10}{3}}H_0^{}+{\displaystyle \frac{4}{3}}H_1^{}{\displaystyle \frac{2}{15}}H_2^{}`$ (248)
in (247). This combination of Fermi-liquid parameters corresponds to a $`\mathrm{}=\mathrm{}^{}=1,J=0`$ distortion of the Fermi sea . We will see later that the result obtained with the naive Landau argument may not be the whole story, since the one-pion-exchange contribution disagrees, though by a small amount, with the chiral Lagrangian prediction derived below.
#### 6.2.2 Chiral Lagrangian prediction
We now calculate the axial charge using our chiral Lagrangian that reproduced the Landau-Migdal formula for the convection current. Consider first only the pion-exchange contribution. In this case we can take the unperturbed nucleon propagator to carry the free space mass $`M`$. The single-particle transition operator corresponding to Fig. 10a is given by
$`A_{0}^{}{}_{1body}{}^{i}=g_A{\displaystyle \frac{๐๐}{M}}{\displaystyle \frac{\tau ^i}{2}}.`$ (249)
There is no contribution of the type of Fig. 10b because of the (G-)parity conservation. The only contribution to the two-body current comes from Fig. 10c and is of the form
$`A_{0}^{}{}_{2body}{}^{i}=g_A{\displaystyle \frac{๐๐}{M}}{\displaystyle \frac{\tau ^i}{2}}\mathrm{\Delta }`$ (250)
with
$`\mathrm{\Delta }={\displaystyle \frac{f^2k_FM}{2g_A^2m_\pi ^2\pi ^2}}\left(I_0I_1{\displaystyle \frac{m_\pi ^2}{2k_F^2}}I_1\right).`$ (251)
The factor $`(1/g_A^2)`$ in (251) arose from replacing $`\frac{1}{f_\pi ^2}`$ by $`\frac{g_{\pi NN}^2}{g_A^2M^2}`$ using the Goldberger-Treiman relation.
Now consider what happens when the vector degrees of freedom are taken into account. Within the approximation adopted, the only thing that needs be done is to implement the BR scaling. The direct intervention of the vector mesons $`\rho `$ and $`\omega `$ in the axial-charge operator is suppressed by the chiral counting, so they will be ignored here. This means that in the single-particle charge operator, all that one has to do is to replace $`M`$ by $`M^{}=M\mathrm{\Phi }`$ in (249):
$`A_{0}^{}{}_{1body}{}^{i}=g_A{\displaystyle \frac{๐๐}{M\mathrm{\Phi }}}{\displaystyle \frac{\tau ^i}{2}}`$ (252)
and that in the two-body charge operator (250), $`f_\pi `$ should be replaced by $`f_\pi \mathrm{\Phi }`$ and $`M`$ by $`M\mathrm{\Phi }`$:
$`A_{0}^{}{}_{2body}{}^{i}=g_A{\displaystyle \frac{๐๐}{M\mathrm{\Phi }}}{\displaystyle \frac{\tau ^i}{2}}\mathrm{\Delta }.`$ (253)
In the two-body operator, there is a factor $`(g_A/f_\pi )`$ coming from the $`\pi NN`$ vertex which as mentioned before, is assumed to be non-scaling at least up to nuclear matter density , in consistency with the observation that the pion-exchange current does not scale in medium.
The total predicted by the chiral Lagrangian (modulo higher-order corrections) is then
$`g_A{\displaystyle \frac{๐๐}{M\mathrm{\Phi }}}{\displaystyle \frac{\tau ^i}{2}}(1+\mathrm{\Delta }).`$ (254)
which differs from the charge operator obtained by the Landau method, (238).
#### 6.2.3 Comparison between vector and axial current
An immediate question (to which we have no convincing answer) is whether or not the difference between the two approaches โ the Fermi-liquid vs. the chiral Lagrangian โ is genuine or a defect in either or both of the approaches. One possible cause of the difference could be that both the assumed localized quasiparticle charge, Eq.(233), and the effective axial charge, Eq.(239), are incomplete. We have looked for possible additional terms that could contribute but we have been unable to find them. So while not ruling out this possibility, we turn to the possibility that the difference is genuine.
It is a well-known fact that the conservation of the vector current assures that the EM charge or the weak vector charge is $`g_V=1`$ but the conservation of the axial vector charge does not constrain the value of the axial charge $`g_A`$ , that is, $`g_A`$ can be anything. This is because the axial symmetry is spontaneously broken. In the Wigner phase in which the axial symmetry would be restored, one would expect that $`g_A=1`$. It therefore seems that the Goldstone structure of the โvacuumโ of the nuclear matter is at the origin of the difference.
To see whether there can be basic differences, let us look at the effect of the pion field. The cancellation between the two-body current $`๐ฑ_{2body}^\pi `$ (211) and $`๐ฑ_{ph}^\pi `$ (220) leaving only a term that changes $`M^{}`$ to $`m_L^{}`$ in the one-body operator with a BR scaling mass, Eq.(210), in the EM case can be understood as follows. Both terms involve the two-body interaction mediated by a pion-exchange. It is obvious how this is so in the latter. To see it in the former, we note that it involves the insertion of an EM current in the propagator of the pion. Thus the sum of the two terms corresponds to the insertion of an EM current in all internal hadronic lines of the one-pion exchange self-energy graph of the nucleon. The two-body pionic current โ that together with the single-particle current preserves gauge invariance โ is in turn related to the one-pion-exchange potential $`V_\pi `$. Therefore what is calculated is essentially an effect of a nuclear force. Now the point is that the density-dependent part of the sum (that is, the ones containing one hole line) โ apart from a term that changes $`M^{}`$ to $`m_L^{}`$ in (210) โ vanishes in the $`\omega /q0`$ limit. In contrast, the cancellation between (247) and (234) in the case of the axial charge, has no corresponding interpretation. While the one-pion-exchange interaction is involved in the particle-hole term (234), (247) cannot be interpreted as an insertion of the axial vector current into the pion propagator since such an insertion is forbidden by parity. In other words, Eq.(247) does not have a corresponding Feynman graph which can be linked to a potential. We interpret this as indicating that there is no corresponding Landau formula for the axial charge in the same sense as in the vector current case.
In a chiral Lagrangian formalism, each term is associated with a Feynman diagram. As mentioned, there is no contribution to the convection current from a diagram of the type Fig. 10c (apart from a gauge non-invariant off-shell term which cancels the counter part in Fig. 10b). Instead this diagram renormalizes the spin gyromagnetic ratio. In contrast, the corresponding diagram for the axial current does contribute to the axial charge (250). As first shown in , the contribution from Fig.10c for both the vector current and the axial-vector current is current algebra in origin and constrained by chiral symmetry. Furthermore it does not have a simple connection to nuclear force. While the convection current is completely constrained by gauge invariance of the EM field, and hence chiral invariance has little to say, both the EM spin current and the axial charge are principally dictated by the chiral symmetry. This again suggests that the Landau approach to the axial charge cannot give the complete answer even at the level of quasiparticle description. There is however a caveat here: in the Landau approach, the nonlocal pionic and local four-Fermion interactions (198) enter together in an intricate way as we saw in the EM case. Perhaps this is also the case in the axial charge, with an added subtlety due to the presence of Goldstone pions. It is possible that the difference is due to the contribution of the four-Fermion interaction term to (248) which cancels out in the limit $`\omega /q0`$ but contributes in the $`q/\omega 0`$ limit. This term cannot be given a simple interpretation in terms of chiral Lagrangians. Amusingly the difference between the results (see below) turns out to be small.
#### 6.2.4 Numerical comparison
To compare the two results, we rewrite the sum of (246) and (247), i.e., โLandau axial chargeโ (LAC), using (41) and (192)
$`A_{0}^{}{}_{LAC}{}^{i}=g_A{\displaystyle \frac{๐๐}{M\mathrm{\Phi }}}{\displaystyle \frac{\tau ^i}{2}}(1+\stackrel{~}{\mathrm{\Delta }})`$ (255)
where
$`\stackrel{~}{\mathrm{\Delta }}={\displaystyle \frac{f^2k_FM\mathrm{\Phi }}{4\pi ^2m_\pi ^2}}\left(I_0I_1+{\displaystyle \frac{3m_\pi ^2}{2k_F^2}}I_1\mathrm{\Phi }^1\right)`$ (256)
and the sum of (252) and (253), i.e., the โcurrent-algebra axial chargeโ(CAAC), as
$`A_{0}^{}{}_{CAAC}{}^{i}=g_A{\displaystyle \frac{๐๐}{M\mathrm{\Phi }}}{\displaystyle \frac{\tau ^i}{2}}(1+\mathrm{\Delta })`$ (257)
where
$`\mathrm{\Delta }={\displaystyle \frac{f^2k_FM}{2g_A^2m_\pi ^2\pi ^2}}\left(I_0I_1{\displaystyle \frac{m_\pi ^2}{2k_F^2}}I_1\right).`$ (258)
We shall compare $`\stackrel{~}{\mathrm{\Delta }}`$ and $`\mathrm{\Delta }`$ for two densities $`\rho =\frac{1}{2}\rho _0`$ ($`k_F=1.50m_\pi `$) and $`\rho =\rho _0`$ ($`k_F=1.89m_\pi `$) where $`\rho _0`$ is the normal nuclear matter density $`0.16/\mathrm{fm}^3`$.
For numerical estimates, we take
$`\mathrm{\Phi }(\rho )=\left(1+0.28{\displaystyle \frac{\rho }{\rho _0}}\right)^1`$ (259)
which gives $`\mathrm{\Phi }(\rho _0)=0.78`$ found in QCD sum rule calculations . Somewhat surprisingly, the resulting values for $`\stackrel{~}{\mathrm{\Delta }}`$ and $`\mathrm{\Delta }`$ are close; they agree within 10%. For instance at $`\rho \rho _0/2`$, $`\stackrel{~}{\mathrm{\Delta }}0.48`$ while $`\mathrm{\Delta }0.43`$ and at $`\rho \rho _0`$, $`\stackrel{~}{\mathrm{\Delta }}0.56`$ while $`\mathrm{\Delta }0.61`$. Whether this close agreement is coincidental or has a deep origin is not known.
#### 6.2.5 Test: axial charge transition in heavy nuclei
The axial charge transition in heavy nuclei
$`A(J^+)B(J^{})`$ (260)
with change of one unit of isospin $`\mathrm{\Delta }T=1`$ provides a test of the axial charge operator (257) or (255). To check this, consider the Warburton ratio $`ฯต_{MEC}`$
$`ฯต_{MEC}=M_{exp}/M_{sp}`$ (261)
where $`M_{exp}`$ is the measured matrix element for the axial charge transition and $`M_{sp}`$ is the theoretical single-particle matrix element. There are theoretical uncertainties in defining the latter, so the ratio is not an unambiguous object but what is significant is Warburtonโs observation that in heavy nuclei, $`ฯต_{MEC}`$ can be as large as 2:
$`ฯต_{MEC}^{HeavyNuclei}=1.92.0.`$ (262)
More recent measurements โ and their analyses โ in different nuclei quantitatively confirm this result of Warburton.
To compare our theoretical prediction with the Warburton analysis, we calculate the same ratio using (257)
$`ฯต_{MEC}^{CAAC}=\mathrm{\Phi }^1(1+\mathrm{\Delta }).`$ (263)
The formula(257) differs from what was obtained in in that here the non-scaling in medium of the pion mass and the ratio $`g_A/f_\pi `$ is taken into account. We believe that the scaling used in (which amounted to having $`\mathrm{\Delta }/\mathrm{\Phi }`$ in place of $`\mathrm{\Delta }`$ in (263)) is not correct.
The enhancement corresponding to the โLandau formulaโ (255) is obtained by replacing $`\mathrm{\Delta }`$ by $`\stackrel{~}{\mathrm{\Delta }}`$ in (263). Using the value for $`\mathrm{\Phi }`$ and $`\mathrm{\Delta }`$ at nuclear matter density, we find
$`ฯต_{MEC}^{th}2.1(2.0)`$ (264)
in good agreement with the experimental results of and . Here the value in parenthesis is obtained with the Landau formula (255). The difference between the two formulas (i.e., current algebra vs. Landau) is indeed small. This is a check of the scaling of $`f_\pi `$ in combination with the scaling of the Gamow-Teller constant $`g_A`$ in medium.
## 7 Summary
An attempt is made and some success is obtained in this review to relate an effective chiral Lagrangian to an effective field theory for nuclear matter. The aim is to bridge between what we know at normal nuclear density and what can be expected under the extreme condition, relevant in neutron stars and in relativistic heavy ion collisions. Furnstahl, Serot and Tangโs effective chiral model Lagrangian FTS1 , which describes successfully the phenomenology of finite nuclei and infinite nuclear matter, is taken to imply that an effective chiral Lagrangian calculated in high chiral orders corresponds to Lynnโs chiral soliton with the chiral liquid structure in mean field. This provides the ground state around which quantum fluctuations can be calculated. Note that FTS1 is simply one of the available theories that are consistent with the symmetries of QCD and successful phenomenologically. We do not imply that FTS1 is the best one can construct as an effective theory of nuclear matter.
The scalar sector in FTS1 develops a large anomalous dimension, which is interpreted as a signal of a strong coupling situation. It is suggested that the strong coupling theory can be transformed into a weak coupling theory if the chiral Lagrangian is rewritten in terms of the parameters given by BR scaling. A simple model, whose mass parameters are BR-scaled, is constructed and is shown to describe ground state properties of nuclear matter very well with fits comparable to the full FTS1 theory. The simple BR-scaled Lagrangian gives the background at any arbitrary density around which fluctuations can be calculated. Tree diagrams yield the dominant contributions. It is shown that we can make simple Walecka-type models, including our simple model. The models are thermodynamically consistent and the dependence of parameters on density are represented by the interactions of hadrons. One can also map the density-dependent model Lagrangian into relativistic Landau Fermi-liquid theory. Thus a quasiparticle picture of a strongly correlated system at densities away from the normal nuclear matter density is obtained.
The BR scaling parameter $`\mathrm{\Phi }`$ has been identified with a Landau Fermi-liquid parameter by means of nuclear responses to the EM convection current. The Landau effective mass of the nucleon $`m_L^{}`$ is given in terms of $`\mathrm{\Phi }`$ and pion cloud, i.e. the Goldstone boson of the broken chiral symmetry, through the Landau parameter $`\stackrel{~}{F}_1^\pi `$. The relation between the exchange current correction to the orbital gyromagnetic ratio $`\delta g_l`$ and $`m_L^{}`$ provides the crucial link between $`\mathrm{\Phi }`$ and $`F_1^\omega `$ which comes from the massive degree of freedom in the isoscalar vector channel dominated by the $`\omega `$ meson. The axial charge transition in heavy nuclei provides a relation between $`\mathrm{\Phi }`$ and the in-medium pion decay constant $`f_\pi ^{}/f_\pi `$. These relations are found to be satisfied very accurately and to connect physics of relativistic heavy ion collision data, e. g., dilepton data of CERES and nucleon and kaon flow data of FOPI(4 pi multiparticle detector) and KaoS(Kaon Spectrometer), etc. to low energy spectroscopic properties, e.g., $`m_L^{}`$, $`\delta g_l`$, etc. in heavy nuclei via BR scaling.
## 8 Open issues
While as an exploration our results are satisfying, there are several crucial links that remain conjectural in the work and require a lot more work. We mention some issues for future studies.
* We have not yet established in a convincing way that a nontopological soliton coming from a high-order effective chiral Lagrangian accurately describes nuclear matter that we know of. The first obstacle here is that a realistic effective Lagrangian that contains sufficiently high-order loop corrections including non-analytic terms has not yet been constructed. Lynnโs argument for the existence of such a soliton solution and identification with a drop of nuclear matter is based on a highly truncated Lagrangian (ignoring non-analytic terms). We are simply assuming that the FTS1 Lagrangian is a sufficiently realistic version (in terms of explicit vector and scalar degrees of freedom that are integrated out by Lynn) of Lynnโs effective Lagrangian. To prove that this assumption is valid is an open problem.
* We do not understand clearly the role and the origin of the anomalous dimension $`d_{an}5/3`$ for the quarkonium scalar field in FTS1. It is an interesting problem how the scalar in FTS1 comes to include higher order interactions in its anomalous scaling dimension through decimations. And our argument for interpreting the FTS1 with such large anomalous dimension as a strong-coupling theory which can be reinterpreted in terms of a weak-coupling theory expressed with BR scaling is heuristic at best and needs to be sharpened, although our results strongly indicate that it is correct.
* There is also the practical question as to how far in density the predictive power of the BR-scaled effective Lagrangian can be pushed. In our simple numerical calculation, we used a parameterization for the scaling function $`\mathrm{\Phi }(\rho )`$ of the simple geometric form which can be valid, if at all, up to the normal matter density as seems to be supported by QCD sum rule and dynamical model calculations. At higher densities, the form used has no reason to be accurate. By using the empirical information coming from nucleon and kaon flows, one could infer its structure up to, say, $`\rho 3\rho _0`$ and if our argument for kaon condensation is correct โ and hence kaon condensation takes place at $`\rho <3\rho _0`$, then this will be good enough to make a prediction for the critical density for kaon condensation. In calculating compact-star properties in supernovae explosions, however, the equation of state for densities considerably higher than the normal matter density, say, $`\rho >5\rho _0`$ is required. It is unlikely that this high density can be accessed within the presently employed approximations. Not only will the structure of the scaling function $`\mathrm{\Phi }`$ be more complicated but also the correlation terms that are small perturbations at normal density may no longer be so at higher densities, as pointed out by Pandharipande, Pethick, and Thorsson who approach the effect of correlations from the high-density limit. In particular, the notion of the scaling function $`\mathrm{\Phi }`$ will have to be modified in such a way that it will become a non-linear function of the fields that figure in the process. This would alter the structure of the Lagrangian field theory. Furthermore there may be a phase transition (such as spontaneously broken Lorentz symmetry, Georgi vector limit, chiral phase transition or meson condensation) lurking nearby in which case the present theory would have already broken down. These caveats will have to be carefully examined before one can extrapolate the notion of BR scaling to a high-density regime as required for a reliable calculation of the compact-star structure. How the scaling parameters extrapolate beyond normal nuclear matter density is not predicted by theory and should be deduced from lattice measurements and heavy-ion experiments that are to come. Corrections to BR scaling as massive mesons approach on-shell need be taken into account. The fit to the available CERES data indicates however that the extrapolation to higher density โ perhaps up to the chiral phase transition โ is at least approximately correct under the conditions that prevail in nucleus-nucleus collisions at SPS energies. How this could come about was discussed in .
* In addition the behavior of $`g_v^{}`$ at $`\rho >\rho _0`$ also deviates from our simple form. It is expected to drop more rapidly . Indeed a recent calculation of kaon attraction to $`๐ช(Q^2)`$ in chiral perturbation theory that is highly constrained by the ensemble of on-shell kaon-nucleon data and that includes both Pauli and short-range correlations for many-body effects is found to give at most about 120 MeV attraction at nuclear matter density. Thus the crucial input here is the strength of the $`K^{}`$-nuclear interaction in dense medium. The attraction decreases from the analysis of the $`K`$-mesic atom by Friedman, Gal, and Batty indicating the 200 MeV attraction. If the attraction came down to $`100120`$ MeV as found in , this would give a strong constraint on the constants that enter in the four-Fermi interactions in the chiral Lagrangian. This would presumably account for the need for a dropping vector coupling $`g_v^{}`$ required for $`\rho >\rho _0`$. Moreover Kim and Lee found recently by renormalization group analysis that the coupling constant $`g_{\rho NN}`$ drops as density increases. This crucial information is also expected to come from on-going heavy-ion experiments.
* Although we show that BR scaling parameters can be written in terms of Landau parameter via nonrelativistic EM current, it remains to formulate the relativistic mapping along the line developed in Section 5.5 where thermodynamic properties of a simple BR-scaled chiral model Lagrangian in the mean field were shown to be consistent with relativistic Landau formula derived in Section 2.3. This work is needed to go to higher density region. Such a work is in progress. Furthermore quasiquarks must become relevant degrees of freedom at high density. Thus quasihadron liquid is shifted to quasiquark liquid as density increases. The investigation of the change in the shift may give a good way to connect the low density physics to the higher one.
* As seen in Section 6.2 it is not clear how we deal with Goldstone bosons in the scheme of Landau-Migdal approach. The Landau axial charge and current algebra axial charge are not the same, though they give similar numerical values. We do not know even whether it is possible or not that the Landau-Migdal approach can treat Goldstone boson properly. The study of it will show the way to treat the Landau Fermi-liquid theory and its scope.
* Finally one could ask more theoretical questions as to in what way our effective Lagrangian approach is connected to standard chiral effective theory, which does not concern scale symmetry and its anomaly, proper and if the theory is to be fully predictive, how one can proceed to calculate the corrections to the tree-diagram results we have obtained. The first issue, a rigorous derivation of BR scaling starting from an effective chiral action via multiple scale decimations required for the problem is yet to be formulated but the main ingredients, both theoretical and phenomenological, seem to be available. The second issue is of course closely tied with what the appropriate expansion parameter is in the theory. These matters are addressed in the paper but they are somewhat scattered all over the place and it might be helpful to summarize them here. The answers to these questions are not straightforward since there are two stages of โdecimationโ in the construction of our effective Lagrangian: The first is the elimination of high-energy degrees of freedom for the effective Lagrangian that gives rise to a soliton (i.e., chiral liquid) and here the relevant scale is the chiral symmetry breaking scale $``$ 1 GeV and the second is that given a chiral liquid which we argued can be identified as the Fermi-liquid fixed point, the decimation involved here is for the excitations of scale $`\mathrm{\Lambda }`$ above (and below) the Fermi surface for which the expansion is made in $`1/N`$. As discussed in Section 2.2, $`1/N`$ $``$ $`\mathrm{\Lambda }/k_F`$ where $`\mathrm{\Lambda }`$ is the cutoff in the Fermi system. In bringing in a BR-scaled chiral Lagrangian, we are relying on chiral symmetry considerations applied to a system with a density defined by nuclear matter. Thus the link to QCD proper of the effective theory we use for describing fluctuations around the nuclear matter ground state must be tenuous at best. As recently re-emphasized by Weinberg , low-energy effective theories need not be in one-to-one correspondence with a fundamental theory meaning that one low-energy effective theory could arise through decimation from several different โfundamentalโ theories. This applies not only to theories with global symmetry but also to those with local gauge symmetry. In the present case, this aspect is more relevant since there is a change in degrees of freedom between the nonperturbative regime in which we are working and the perturbative regime in which QCD proper is operative.
## Acknowledgements
I am very grateful to Professor D.-P. Min and Professor M. Rho for their guidance and encouragement throughout my graduate years. Collaboration with Professors G.E. Brown and B. Friman has been a great pleasure to me. I would like to thank C.-H. Lee and R. Rapp for useful comments and discussions. This work was supported in part by the U.S. Department of Energy under DE-FG02-88ER40388, by the Korea Science and Engineering Foundation through Center for Theoretical Physics of Seoul National University, and by the Korea Ministry of Education under BSRI-98-2418.
## Appendix
## A: Effect of many-body correlations on EOS
In this appendix, we briefly discuss the sensitivity of the EOS to the correlation parameters of (5.4) at a density $`\rho >\rho _0`$. This is shown in Fig. A.1. While the parameter sets B1, B2, B3 and B4 give more or less the same equilibrium density and binding energy (see Table 4), the parameter set B2 has an instability and B4 a local minimum at $``$ 2 times the normal matter density whereas the sets B1 and B3 give a stable state at all density, possibly up to meson condensations and/or chiral phase transition. It is not clear what this means for describing fluctuations at a density above $`\rho _0`$ but it indicates that given data at ordinary nuclear matter density, it will not be feasible to extrapolate in a unique way to higher densities unless one has constraints from experimental data at the corresponding density. In our discussion, we relied on the data from KaoS and FOPI collaborations to avoid the fine-tuning of the parameters.
## B: Relativistic calculation of $`F_1^\pi `$
In the text, the Landau parameter $`F_1^\pi `$ (or $`f^\pi `$) was calculated nonrelativistically via the Fock term of Fig. B.1. Here we calculate it relativistically by Fierz-transforming the one-pion-exchange graph and taking the Hartree term. This procedure is important for implementing relativity in the connection between Fermi-liquid theory and chiral Lagrangian theory along the line discussed by Baym and Chin .
The one-pion-exchange potential in Fig. B.1 is
$`V_\pi =g_{\pi NN}^2(๐_{21}๐_{43}){\displaystyle \frac{\overline{u}_2\gamma ^5u_1\overline{u}_4\gamma ^5u_3}{(p_2p_1)^2m_\pi ^2}}.`$ (B.1)
The Dirac spinors are normalized by
$`u^{}(p,s)u(p,s^{})=\delta _{ss^{}}.`$ (B.2)
By a Fierz transformation, we have
$`๐_{21}๐_{43}={\displaystyle \frac{1}{2}}(3\delta _{41}\delta _{23}๐_{41}๐_{32})`$ (B.3)
and
$`\overline{u}_2\gamma ^5u_1\overline{u}_4\gamma ^5u_3`$ $`=`$ $`{\displaystyle \frac{1}{4}}[\overline{u}_4u_1\overline{u}_2u_3\overline{u}_4\gamma ^\mu u_1\overline{u}_2\gamma _\mu u_3`$
$`+\overline{u}_4\sigma ^{\mu \nu }u_1\overline{u}_2\sigma _{\mu \nu }u_3+\overline{u}_4\gamma ^\mu \gamma ^5u_1\overline{u}_2\gamma _\mu \gamma ^5u_3+\overline{u}_4\gamma ^5u_1\overline{u}_2\gamma ^5u_3].`$
Remembering a minus sign for the fermion exchange, we obtain the corresponding pionic contribution to the quasiparticle interaction at the Fermi surface, $`f^\pi =V_\pi (๐_1=๐_4=๐,๐_2=๐_3=๐^{},๐^2=๐^2=k_F^2)`$ (see (2)). Decomposing $`f^\pi `$ as
$`f^\pi ={\displaystyle \frac{3๐๐^{}}{2}}(f_S+f_V+f_T+f_A+f_P)`$ (B.5)
where $`S`$, $`V`$, $`T`$, $`A`$ and $`P`$ represent scalar, vector, tensor, axial vector and pseudoscalar channel respectively, we find
$`f_S`$ $`=`$ $`{\displaystyle \frac{M^4f^2}{E_F^2m_\pi ^2}}{\displaystyle \frac{1}{q^2+m_\pi ^2}}`$
$`f_V`$ $`=`$ $`{\displaystyle \frac{M^4f^2}{E_F^2m_\pi ^2}}{\displaystyle \frac{1}{q^2+m_\pi ^2}}\left(1+{\displaystyle \frac{q^2}{2M^2}}\right)`$
$`f_T`$ $`=`$ $`{\displaystyle \frac{M^4f^2}{E_F^2m_\pi ^2}}{\displaystyle \frac{1}{q^2+m_\pi ^2}}\left(๐๐^{}(1+{\displaystyle \frac{q^2}{2M^2}})+{\displaystyle \frac{2๐^{}๐๐๐^{}๐๐๐^{}๐๐๐^{}๐^{}๐^{}}{2M^2}}\right)`$
$`f_A`$ $`=`$ $`{\displaystyle \frac{M^4f^2}{E_F^2m_\pi ^2}}{\displaystyle \frac{1}{q^2+m_\pi ^2}}\left(๐๐^{}{\displaystyle \frac{2๐๐๐^{}๐^{}๐๐๐^{}๐๐๐^{}๐^{}๐^{}}{2M^2}}\right)`$
$`f_P`$ $`=`$ $`0.`$ (B.6)
with $`E_F=\sqrt{k_F^2+M^2}`$ and $`q=|๐๐^{}|`$. Thus we obtain
$`f^\pi `$ $`=`$ $`{\displaystyle \frac{f^2}{m_\pi ^2}}{\displaystyle \frac{M^2}{E_F^2}}{\displaystyle \frac{1}{q^2+m_\pi ^2}}\left(๐๐๐^{}๐{\displaystyle \frac{q^2(1๐๐^{})}{2}}\right){\displaystyle \frac{3๐๐^{}}{2}}`$
$`=`$ $`{\displaystyle \frac{1}{3}}{\displaystyle \frac{f^2}{m_\pi ^2}}{\displaystyle \frac{M^2}{E_F^2}}{\displaystyle \frac{q^2}{q^2+m_\pi ^2}}\left(3{\displaystyle \frac{๐๐๐^{}๐}{q^2}}๐๐^{}+{\displaystyle \frac{1}{2}}(3๐๐^{})\right){\displaystyle \frac{3๐๐^{}}{2}}.`$
In the nonrelativistic limit, $`E_FM`$ and we recover (191). The factor $`M/E_F`$ comes since there is one particle in the unit volume which decreases relativistically as the speed increases. Note that only $`f_S`$ and $`f_V`$ in (B.6) are spin-independent and contribute to $`F_1^\pi `$. The $`f_S`$ is completely canceled by the leading term of $`f_V`$ with the remainder giving $`F_1^\pi `$. In this way of deriving the Landau parameter $`F_1`$, it is the vector channel that plays the essential role.
## C: Relativistic calculation of $`F_1(\omega )`$ and $`๐ฑ_{2body}^\omega `$
Here we compute the contribution of vector meson channel to Landau parameter $`F_1`$ and EM current relativistically. The way to compute the contribution of $`\rho `$-meson channel is almost the same as $`\omega `$ channel. So we treat here $`\omega `$-meson channel only.
The blob in Fig. 11a corresponding to four-Fermi interactions can be expanded as Fig. C.1 in random phase approximation.
One bubble is represented by
$`\mathrm{\Pi }^{\mu \nu }=\underset{\omega _q0}{lim}\underset{q0}{lim}2{\displaystyle \frac{d^4p}{(2\pi )^4}\mathrm{tr}[\gamma ^\mu S(p)\gamma ^\nu S(p+q)]}`$ (C.1)
where $`S(p)`$ is the Fermion propagator. The factor 2 come from isospin contribution. In the presence of Fermi sea, we can divide $`S(p)`$ into the four parts;
$`S(p)`$ $`=`$ $`{\displaystyle \frac{1}{2\omega _p}}[(\omega _p\gamma _0๐๐ธ+M^{})({\displaystyle \frac{1n_p}{p_0\omega _p+i\delta }}+{\displaystyle \frac{n_p}{p_0\omega _pi\delta }})`$ (C.2)
$`+{\displaystyle \frac{\omega _p\gamma _0+๐๐ธM^{}}{p_0+\omega _pi\delta }}]`$
with $`\omega _p=\sqrt{M^2+p^2}`$ and $`n_p=\theta (k_Fp)`$ at $`T=0`$. The first term is the free particle propagator in vacuum. The second is the particle propagator for Pauli-blocked state in medium. The third is for hole and the fourth is for antiparticle. Since vacuum contribution, i.e. antiparticle-particle in vacuum contribution, is canceled by counter terms and particle-hole contribution vanishes in our limit $`q/\omega _q0`$, the antiparticle-particle in Pauli-blocked state contribution remains. Then (C.1) becomes
$`\mathrm{\Pi }^{\mu \nu }=`$ $`{\displaystyle }{\displaystyle \frac{d^3p}{(2\pi )^3}}{\displaystyle \frac{n_p}{4\omega _p^3}}\mathrm{tr}[\gamma ^\mu (\omega _p\gamma _0๐๐ธ+M^{})\gamma ^\nu (\omega _p\gamma _0+๐๐ธM^{})`$ (C.3)
$`+`$ $`\gamma ^\mu (\omega _p\gamma _0+๐๐ธM^{})\gamma ^\nu (\omega _p\gamma _0๐๐ธ+M^{})].`$
Because of rotational invariance and zero energy-momentum transfer, only $`\mathrm{\Pi }^{ii}`$ does not vanish.
$`\mathrm{\Pi }^{ii}={\displaystyle \frac{4}{3}}{\displaystyle ^{k_F}}{\displaystyle \frac{d^3p}{(2\pi )^3}}{\displaystyle \frac{3M^2+2p^2}{\omega _p^3}}={\displaystyle \frac{\rho }{E_F}}`$ (C.4)
with $`E_F=\sqrt{M^2+k_F^2}`$.
The quasiparticle interaction in $`\omega `$-meson channel in Fig. C.1 gives
$`f_{pp^{}}^\omega `$ $`=`$ $`C_\omega ^2\left[\overline{u}(p)\gamma ^\mu u(p)\overline{u}(p^{})\gamma _\mu u(p^{})\overline{u}(p)\gamma _iu(p)\left({\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}(C_\omega ^2{\displaystyle \frac{\rho }{E_F}})^i\right)\overline{u}(p^{})\gamma ^iu(p^{})\right]`$ (C.5)
$`=`$ $`C_\omega ^2\left(\overline{u}(p)\gamma ^\mu u(p)\overline{u}(p^{})\gamma _\mu u(p^{})+\overline{u}(p)\gamma _iu(p){\displaystyle \frac{C_\omega ^2\rho /E_F}{1+C_\omega ^2\rho /E_F}}\overline{u}(p^{})\gamma ^iu(p^{})\right)`$
$`=`$ $`C_\omega ^2C_\omega ^2{\displaystyle \frac{๐๐^{}}{\mu E_F}}`$
with chemical potential $`\mu =E_F+C_\omega ^2\rho `$. Thus
$`F_1(\omega )=C_\omega ^2{\displaystyle \frac{2k_F^3}{\pi ^2\mu }}.`$ (C.6)
And the EM current in Fig. 11a is
$`๐ฑ_{2body}^\omega `$ $`=`$ $`\overline{u}(k)๐ธu(k){\displaystyle \frac{C_\omega ^2\rho /E_F}{1+C_\omega ^2\rho /E_F}}`$ (C.7)
$`=`$ $`{\displaystyle \frac{๐}{\mu }}{\displaystyle \frac{\stackrel{~}{F}_1(\omega )}{6}}.`$
|
warning/0006/hep-th0006052.html
|
ar5iv
|
text
|
# References
June 2000
OU-HET 351
Description of Intersecting Branes
via Tachyon Condensation
Takao Suyama <sup>1</sup><sup>1</sup>1e-mail address : suyama@funpth.phys.sci.osaka-u.ac.jp
Department of Physics, Graduate School of Science, Osaka University,
Toyonaka, Osaka, 560-0043, Japan
Abstract
We construct a model describing BPS brane-systems using low energy effective theory of brane-antibrane system. Both parallel branes and intersecting branes can be treated by this model. After tachyon condensation, the dynamics of fluctuations around such brane-systems is supersymmetric if the degrees of freedom are restricted on the branes. The form of the tachyon potential and the application of this model to the black hole physics are discussed.
1. Introduction
Low energy dynamics of the combined D-brane systems is important, for example, for black hole physics . It is easy to describe such system which contains only one kind of D-branes. The effective theory of the system is the dimensional reduction of the ten-dimensional super Yang-Mills theory . However, construction of effective theory is not so straightforward when the system consists of a combination of D-branes with different dimensions. In most cases, one pays attention to a specific part of the worldvolume and consider the effective theory living only on that region . For example, the effective theory of the D1-D5 system is a two-dimensional superconformal field theory living on the D1-branes. Although this is enough to discuss the physics of the black holes, it is preferable to have another way of describing the system as a whole.
The difficulty in such description may be how to treat the brane-intersections. When one writes down the effective theory of the massless modes on the whole brane-system, the theory contains a delta-function-like interaction at the intersection. It would not be an easy task to deal with such theory, although this description has some applications .
In this paper we propose another way of describing the D-brane system which includes intersecting branes. Our idea is to apply the tachyon condensation . In this situation, pair annihilations of D-branes can be discussed and in some cases the lower dimensional branes can remain as a result. The worldvolume of the resulting brane is determined by the form of the tachyon field and, in particular, it need not be a smooth manifold. Therefore the theory describing the above phenomenon would be able to handle the low energy dynamics of the intersecting branes.
This paper is organized as follows. We review the story of the tachyon condensation briefly in section 2. The construction of the model is discussed in section 3. This model can describe both parallel branes (section 4) and intersecting branes (section 5). Section 6 is devoted to discussions. In Appendix, we summarize the properties of the vortex solutions used in this paper.
2. Review of the tachyon condensation
In this section, we will review some properties of Dp-$`\overline{\text{D}}`$p system. While both branes preserve half of the supersymmetries, this system breaks all of the supersymmetries when they exist together.
There exists a tachyonic mode coming from the string stretched between Dp-brane and $`\overline{\text{D}}`$p-brane, which survives the GSO projection . This indicates the instability of the system. Assuming that the tachyon potential has a minimum, the tachyon would condense and this system decays to some stable vacuum.
Let $`T`$ denote the tachyon field whose potential $`V(T)`$ has a minimum at $`T=T_0`$. When $`T=T_0`$ everywhere, there remains nothing : pair annihilation of the D-branes occurs. This has been checked by examining that $`V(T_0)`$ would exactly cancel the tension of the D-branes, using string field theory . On the other hand, when $`T`$ has some nontrivial form, there remains a lower dimensional brane which is stable but not necessarily BPS . In this case, the tension of the originally unstable brane-antibrane system is cancelled in almost all regions except at those places where the tachyon field behaves nontrivially. The energy therefore remains nonzero only on that regions, and this is regarded as a lower dimensional brane. It is argued that K-theory determines what kind of branes can appear after tachyon condensation .
For example, consider the D1-$`\overline{\text{D}}`$1 system extended along $`x^1`$ direction . Suppose that $`T`$ supports a kink,
$$T(x)=\{\begin{array}{c}+T_0(x^1+\mathrm{})\hfill \\ T_0(x^1\mathrm{})\hfill \end{array}$$
$$T(x^1=0)=0,$$
where we have assumed that $`V(T)=V(T)`$ and $`T=T_0`$ is also a minimum of $`V(T)`$. Then the remaining energy is localized around $`x^1=0`$, which can be regarded as D0-brane. In ordinary Type IIB theory this is of course unstable, but this is stable in Type I theory .
The brane annihilation summarized above is also described in terms of gauge theory on the D-brane worldvolume . The worldvolume theory would be a gauge theory coupled to the tachyon. When the tachyon acquires VEV, the gauge symmetry is broken down by the Higgs mechanism. Thus the gauge fields acquire masses and are decoupled from the low energy effective theory. This indicates the disappearance of the D-branes. The case of nontrivial $`T`$ can be described similarly. If $`T=0`$ in some region, the gauge symmetry is recovered there, and this corresponds to the resulting lower dimensional brane.
In the gauge theory language, the classification of the remaining branes can be done in terms of homotopy groups of the vacuum manifold .
3. The model
We will consider in this section the model of D9-$`\overline{\text{D}}`$9 system in Type IIB theory. The model of Dp-$`\overline{\text{D}}`$p system can be obtained by dimensional reduction.
There is a $`U(1)`$ gauge multiplet on each brane and also a tachyon field $`T`$ which couples to both gauge fields. One linear combination of these gauge fields (so called diagonal $`U(1)`$) will decouple from the dynamics (it is argued that this gauge field becomes a fundamental string , and its generalizations are discussed ). Then the resulting model is ten-dimensional $`U(1)`$ gauge theory coupled to the tachyon field,
$$S=d^{10}x\left\{\frac{1}{4}F_{\mu \nu }F^{\mu \nu }+\frac{i}{2}\overline{\psi }\mathrm{\Gamma }^\mu _\mu \psi |D_\mu T|^2V(T)\right\}$$
(1)
where $`D_\mu T=_\mu T+iA_\mu T`$. $`\mathrm{\Gamma }^\mu `$ are the ten-dimensional gamma matrices and $`\psi `$ is a ten-dimensional Majorana-Weyl spinor. Although there exists a charged massless fermion coming from the string stretched between branes, it is not important for the following discussions and we will ignore it. This action would be a first few terms in the expansion of the more complicated DBI-like action. The spacetime action of non-BPS brane is argued in . It is determined by requiring that it reduces to the ordinary DBI action if the tachyon and half of the massless fermions which are absent on the BPS brane are set to zero. This seems to imply that in the Dp-$`\overline{\text{D}}`$p case the action is also supersymmetric if the tachyon field $`T`$ is set to zero. The action of this system is discussed in .
The action (1) is obviously not supersymmetric. According to the discussion in string theory, however, classical solutions should exist which preserve some of the supersymmetries. The first two terms are invariant under the ordinary transformations.
$`\delta A_\mu ={\displaystyle \frac{i}{2}}\overline{ฯต}\mathrm{\Gamma }_\mu \psi `$ (2)
$`\delta \psi ={\displaystyle \frac{1}{4}}F_{\mu \nu }\mathrm{\Gamma }^{\mu \nu }ฯต`$ (3)
The tachyon should be invariant : $`\delta T=0`$. Then,
$$\delta S=d^{10}x\frac{i}{2}\overline{ฯต}\mathrm{\Gamma }_\mu \psi (iT^{}D^\mu TiD^\mu T^{}T).$$
(4)
This vanishes when $`D_\mu T=0`$ which leads to trivial solutions.
Now we will modify the transformations of the fermions (3) as follows,
$$\delta \psi =\frac{1}{4}F_{\mu \nu }\mathrm{\Gamma }^{\mu \nu }ฯต+f(|T|^2)ฯต,$$
(5)
where $`f(x)`$ is a function which may have spinor indices. Then the variation of the action is
$$\delta S=d^{10}x\frac{i}{2}\overline{\psi }\mathrm{\Gamma }_\mu (iT^{}D^\mu T+iD^\mu T^{}T+2_\mu f(|T|^2))ฯต.$$
(6)
We will show below that for a suitable choice of $`f(x)`$, there exist classical solutions which preserve some of supersymmetries.
4. Parallel D-branes
We will consider for example D4-$`\overline{\text{D}}`$4 system which decays to D2-branes. Suppose that D4-brane and $`\overline{\text{D}}`$4-brane extend along the (01234) directions, and the resulting D2-branes extend along (034) directions. The action of the model is eq.(1) dimensionally reduced to five dimensions. For the solutions corresponding to such branes to exist, the appropriate choice of $`f(x)`$ is
$$f(|T|^2)=\frac{1}{2}s(|T|^2\zeta )\mathrm{\Gamma }^1\mathrm{\Gamma }^2(\zeta >0).$$
(7)
The variation of the action is
$`\delta S={\displaystyle }d^5x[{\displaystyle \frac{i}{2}}\overline{\psi }T^{}(\mathrm{\Gamma }^1(iD_1T+sD_2T)+\mathrm{\Gamma }^2(iD_2TsD_1T)`$ (8)
$`+D_iT(i\mathrm{\Gamma }^i+\mathrm{\Gamma }^{i12})+iA_kT(i\mathrm{\Gamma }^k+\mathrm{\Gamma }^{k12}))ฯต+(h.c.)],`$
where $`i=0,3,4`$ and $`k=5,6,\mathrm{},9`$. In this case $`A_k`$ is the neutral scalars. Thus the conditions for $`\delta S=0`$ are
$`D_1TisD_2T=0`$ (9)
$`D_iT=0`$ (10)
$`A_kT=0,`$ (11)
We assume translational invariance along $`i`$-th direction, thus eq.(10) sets $`A_i=0`$, and also $`A_k=0`$ by eq.(11). $`\delta \psi =0`$ means
$$F_{12}+s(|T|^2\zeta )=0.$$
(12)
We will investigate the stability of this BPS solutions. Their energy is rewritten as follows.
$$E=d^4x\left\{\frac{1}{2}\left(F_{12}+s(|T|^2\zeta )\right)^2+|D_1TisD_2T|^2+s\zeta F_{12}+V(T)\frac{1}{2}(|T|^2\zeta )^2\right\}$$
(13)
If the tachyon potential takes the form
$$V(T)=\frac{1}{2}(|T|^2\zeta )^2,$$
(14)
then eq.(13) becomes
$$E=d^4x\left\{\frac{1}{2}\left(F_{12}+s(|T|^2\zeta )\right)^2+|D_1TisD_2T|^2+s\zeta F_{12}\right\}.$$
(15)
Therefore the BPS solutions are stable topologically.
Now we have the following BPS equations whose solutions preserve 16 supercharges.
$`D_1TisD_2T=0`$ (16)
$`F_{12}+s(|T|^2\zeta )=0`$ (17)
This is the equations for the Nielsen-Olesen vortex . The properties of this solutions are well-known. Some of these are collected in the Appendix. The solutions are labeled by an integer $`n`$ (quantized magnetic flux), and is determined by specifying $`n`$ distinct points in 1-2 plane, at which $`T=0`$. As explained in section 2, zero loci of the tachyon field $`T`$ correspond to the D-brane worldvolume. Thus the $`n`$-vortex solution describes $`n`$ parallel D2-branes.
The worldvolume theory on the D2-branes should be a supersymmetric one. For this, eqs.(9)(10)(11) has to be satisfied even when the fluctuations are included. Let $`a_\mu ,\phi `$ denote the fluctuations of $`A_\mu ,\psi `$ around the vortex solution respectively, then eqs.(9)(10)(11) means
$$a_\mu 0\text{only at}T=0$$
(18)
From the transformations (2), $`\phi `$ is also restricted to the region specified by $`T=0`$. Then the resulting model is supersymmetric and the physical degrees of freedom exist only at the cores of the vortices, as is expected from the D-brane interpretation.
It is interesting that we can construct $`n`$ D-brane system from $`U(1)`$ gauge theory. Moreover system with diferent number of D-branes merely corresponds to taking another classical solution in the same model. In the ordinary approach to the D-brane worldvolume theory, the number of D-branes must be fixed to construct the theory. It might be possible to relate this feature of our model to the second quantization of D-branes.
5. Intersecting branes
We will show in this section that our model can also describe intersecting branes in a similar way which was discussed in the previous section. We will consider the D4-$`\overline{\text{D}}`$4 system which extends along (01234) directions. The action of the model is again eq.(1) dimensionally reduced to five dimensions, and the tachyon potential and supersymmetry transformations are given as eqs.(14)(5)(7). Suppose that this system decays to the intersecting D2-D2โ system, in which D2-branes extend alnog (034) directions and D2โ-branes along (012) directions. The conditions for the remaining supersymmetries are
$`\mathrm{\Gamma }^1\mathrm{\Gamma }^2ฯต=s^{}\mathrm{\Gamma }^3\mathrm{\Gamma }^4ฯต`$ (19)
$`(s^{}=\pm 1).`$
The BPS conditions are then
$`D_1TisD_2T=0`$ (20)
$`D_3Tiss^{}D_4T=0`$ (21)
$`F_{12}+s^{}F_{34}+s(|T|^2\zeta )=0`$ (22)
$`F_{13}s^{}F_{24}=0`$ (23)
$`F_{14}+s^{}F_{23}=0,`$ (24)
where we have assumed $`A_k=0(k=0,5,6,\mathrm{},9)`$ and that the solutions are static. It is easy to check that the solutions of these equations are stable topologically.
At first sight, these conditions are overdetermined. However the last two equations (23)(24) are in fact redundant. From eqs.(20)(21), $`T`$ and $`A_i(i=1,2,3,4)`$ are written as follows.
$`T=\rho ^{\frac{1}{2}}e^{i\omega }`$ (25)
$`A_1={\displaystyle \frac{1}{2}}s_2\mathrm{log}\rho _1\omega `$ (26)
$`A_2={\displaystyle \frac{1}{2}}s_1\mathrm{log}\rho _2\omega `$ (27)
$`A_3={\displaystyle \frac{1}{2}}ss^{}_4\mathrm{log}\rho _3\omega `$ (28)
$`A_4={\displaystyle \frac{1}{2}}ss^{}_3\mathrm{log}\rho _4\omega ,`$ (29)
These solve eqs.(23)(24) automatically. The only nontrivial equation (22) is then
$$\frac{1}{2}\underset{i=1}{\overset{4}{}}(_i)^2\mathrm{log}\rho +\rho \zeta =0.$$
(30)
$`\omega `$ is determind by requiring the regularity of $`A_i`$ and the single-valuedness of $`T`$.
The solutions we would like to find are symmetric under the rotations in 1-2 and 3-4 planes (D2(D2โ)-branes coincide). Now we take the polar coordinates for these planes,
$$(x_1,x_2)(r_1,\theta _1),(x_3,x_4)(r_2,\theta _2),$$
and assume that $`\rho `$ is independent of both $`\theta _1`$ and $`\theta _2`$. Then (30) becomes
$$\left[\frac{^2}{r_1^2}+\frac{1}{r_1}\frac{}{r_1}+\frac{^2}{r_2^2}+\frac{1}{r_2}\frac{}{r_2}\right]\mathrm{log}\rho =2(\rho \zeta ).$$
(31)
The region around $`r_1=r_2=0`$, $`\rho `$ behaves as
$$\rho r_1^{2n_1}r_2^{2n_2}.$$
(32)
For the regularity and the single-valuedness of the solution, $`n_1`$ and $`n_2`$ are positive integers.
It is convenient to make the following change of variables : $`r_1=e^{t_1},r_2=e^{t_2}`$. Then
$$\left[e^{2t_1}\frac{^2}{t_1^2}+e^{2t_2}\frac{^2}{t_2^2}\right]\mathrm{log}\rho =2(\rho \zeta ).$$
(33)
When $`t_2\mathrm{}`$, the second term of LHS is negligible and $`\rho `$ becomes $`t_2`$ (thus $`r_2`$) independent. This means that in this region $`\rho `$ approaches the vortex solution with magnetic flux $`n_1`$ discussed in the previous section. The same is true with the behavior of $`\rho `$ when $`t_1\mathrm{}(r_1\mathrm{})`$ and its flux is $`n_2`$. By the continuity of $`\omega `$, $`n_1`$ and $`n_2`$ take the same values as the ones in eq.(32).
The global behavior of the solutions can also be discussed. Now take $`u=\mathrm{log}\frac{\rho }{\zeta }`$ and define $`๐_+`$ to be the region where $`u>0`$. $`๐_+`$ is a finite region in $`๐^2`$. Then using eq.(33),
$`0`$ $`=`$ $`{\displaystyle _{๐_+}}๐t_1๐t_2\left[u\{e^{2t_2}_1^2ue^{2t_1}_2^2+2\zeta e^{2t_1+2t_2}(e^u1)\}\right]`$ (34)
$`=`$ $`{\displaystyle _{๐_+}}๐t_1๐t_2\left[e^{2t_2}(_1u)^2+e^{2t_1}(_2u)^2+2\zeta e^{2t_1+2t_2}u(e^u1)\right],`$
where we have used that $`u=0`$ at the boundary of $`๐_+`$. The integrand is strictly positive, and therefore, eq.(34) means $`u0`$ (i.e. $`\rho \zeta `$) everywhere.
The above analyses imply that $`\rho =0`$ at $`r_1=0`$ and/or $`r_2=0`$, and $`\rho \zeta `$ away from the cores of the vortices. $`\omega `$ is determined to be $`n_1\theta _1+n_2\theta _2`$, and this gives
$`{\displaystyle \frac{1}{2\pi }}{\displaystyle ๐x_1๐x_2F_{12}}`$ $`=`$ $`n_1`$ (35)
$`{\displaystyle \frac{1}{2\pi }}{\displaystyle ๐x_3๐x_4F_{34}}`$ $`=`$ $`n_2.`$ (36)
Therefore this solution corresponds to $`n_1`$ coincident D2-branes and $`n_2`$ coincident D2โ- branes.
The dynamics of the fluctuations around this solution is supersymmetric if they are restricted at the cores of the vortices, as in the case of parallel D-branes. This means that we can, in principle, construct the effective theory of D-brane system whose worldvolume is not a smooth manifold.
The intersecting D4-D4โ-D4โ system can also be described, starting from D6-$`\overline{\text{D}}`$6 system. The BPS conditions for this case are
$`D_1TisD_2T=0`$
$`D_3Tiss^{}D_4T=0`$
$`D_5Tiss^{\prime \prime }D_6T=0`$
$`F_{12}+s^{}F_{34}+s^{\prime \prime }F_{56}+s(|T|^2\zeta )=0`$
$`F_{13}s^{}F_{24}=0`$
$`F_{14}+s^{}F_{23}=0`$
$`F_{15}s^{\prime \prime }F_{26}=0`$
$`F_{16}+s^{\prime \prime }F_{25}=0`$
$`s^{}F_{35}s^{\prime \prime }F_{46}=0`$
$`s^{}F_{36}+s^{\prime \prime }F_{45}=0`$
$`(s^{},s^{\prime \prime }=\pm 1).`$
As in the D2-D2โ case, the last six equations are redundant and the only nontrivial equation is rewritten as follows,
$$\frac{1}{2}\underset{k=1}{\overset{6}{}}(_k)^2\mathrm{log}\rho +\rho \zeta =0,$$
where $`T=\rho ^{\frac{1}{2}}e^{i\omega }`$. The analysis of the solutions can be done similarly as for the previous case.
6. Discussions
We have discussed the discription of the BPS brane-system via the tachyon condensation. The model (1) can describe both parallel branes and intersecting branes. Moreover the system with different number of D-branes can be treated in the same model. The dynamics of the fluctuations which localize on the branes is supersymmetric, and this model can, in principle, provide a way of describing the effective theory of the brane-system whose worldvolume is not a smooth manifold.
For the BPS D-branes to exist, we have seen that the specific form of the tachyon potential is needed. The existence of BPS D-branes would also require similar restriction on the tachyon potential in the case of more complicated DBI-like action . This might provide some information of the profile of the tachyon potential.
D4-D4โ-D4โ system is considered in the end of the previous section. To this system, one can add D0-branes without breaking any supersymmetries. The bound states of such a system was discussed in and it is conjectured that this has four bosonic states and four fermionic states (although the system considered in is the D-branes wrapped on some cycle in a Calabi-Yau manifold). If the model (1) is generalized to the non-Abelian gauge thoery, there would exist a solution which contains the D0-branes in addition to the D4-branes. Therefore, the above conjecture could be checked by quantizing the collective coordinates of such a solution.
As shown in section 3, $`n`$ parallel D-branes are described in terms of $`U(1)`$ gauge theory. The generic vortex solution corresponds to the separated D-branes and the gauge symmetry on each worldvolume is $`U(1)`$. It should be expected that when two or more vortices coincide, the gauge symmetry is enhanced. This will certainly be achieved in the non-Abelian model. This gauge enhancement might occur already in the $`U(1)`$ model, however.
Acknowledgements
I would like to thank H.Itoyama, T.Matsuo, K.Murakami for valuable discussions. This work is supported in part by JSPS Research Fellowships.
Appendix : Vortex solutions
The Nielsen-Olesen vortex solution is the solution of the following equations ,
$`D_1TiD_2T=0`$ (37)
$`F_{12}+|T|^2\zeta =0,`$ (38)
where $`D_kT=_kT+iA_kT(k=1,2)`$. $`A_k`$ and $`T`$ can be written as
$`T=\rho ^{\frac{1}{2}}e^{i\omega }`$ (39)
$`A_k={\displaystyle \frac{1}{2}}\epsilon _{kl}_l\mathrm{log}\rho _k\omega .`$ (40)
$`\rho `$ is determined by
$$\frac{1}{2}(_1^2+_2^2)\mathrm{log}\rho +\rho \zeta =0.$$
(41)
We now consider the radially symmetric solutions. Taking the polar coordinate $`(r,\theta )`$ and assuming that $`\rho `$ is independent of $`\theta `$, (41) becomes
$$\frac{1}{2}\left(\frac{d^2}{dr^2}+\frac{1}{r}\frac{d}{dr}\right)\mathrm{log}\rho +\rho \zeta =0.$$
(42)
$`\rho `$ is a monotonic function and behaves as
$$\rho \{\begin{array}{c}r^n(r0)\hfill \\ \zeta (r\mathrm{})\hfill \end{array},$$
where $`n`$ is a positive integer. This behavior makes $`A_k`$ singular at $`r=0`$. To eliminate this singularity,
$$\omega =\frac{n}{2}\theta .$$
(43)
For the single-valuedness of $`T`$, $`n`$ must be an even integer.
General solutions are also known . The $`n`$-vortex solution is determined by specifying the positions of $`n`$ points in the plane, at which $`T=0`$ (cores of the vortices).
|
warning/0006/hep-ph0006019.html
|
ar5iv
|
text
|
# References
TU/00/592
KEK-TH-697
Dynamics of barrier penetration in thermal medium: exact result for inverted harmonic oscillator
Sh. Matsumoto<sup>1</sup> and M. Yoshimura<sup>2</sup>
<sup>1</sup> Theory Group, KEK
Oho 1-1 Tsukuba Ibaraki 305-0801 Japan
<sup>2</sup> Department of Physics, Tohoku University
Sendai 980-8578 Japan
ABSTRACT
Time evolution of quantum tunneling is studied when the tunneling system is immersed in thermal medium. We analyze in detail the behavior of the system after integrating out the environment. Exact result for the inverted harmonic oscillator of the tunneling potential is derived and the barrier penetration factor is explicitly worked out as a function of time. Quantum mechanical formula without environment is modifed both by the potential renormalization effect and by a dynamical factor which may appreciably differ from the previously obtained one in the time range of 1/(curvature at the top of potential barrier).
I Introduction
Tunneling phenomena in thermal medium are of both theoretical and practical interest in many areas of physics. For instance, in cosmology there are a variety of tunneling phenomena that may have occured in the evolution of our universe. If the phase transition of the electroweak gauge symmetry is of the first order as assumed in the electroweak scenario of baryogenesis , then the tunneling from a metastable state to the true ground state of the electroweak theory must go through either quantum effect or thermally activated barrier crossing. Another possibility of the first order phase transition in cosmology is the quark-hadron phase transition. We should neither fail to mention the classic example of tunneling phenomenon that takes place in the central core of stars: thermonuclear reaction .
A common problem to all these is the presence of environment: the tunneling we are interested in does not take place for the system in isolation. Under this circumstance we are very much interested in how substantially quantum mechanical formula for the tunneling rate is modified by dissipative interaction with surrounding medium. Despite of this obvious interest, many past works in cosmology and astropysics have relied on simple methods to deal with the tunneling problem, either by using the bounce solution in the Euclidean approach , , or by exploiting some variant of the classical Vant Hoff-Arrhenius law , .
On the other hand, since the pioneering work on quantum dissipation to tunneling phenomena at zero temperature, there have appeared many extended works in condensed matter or statistical physics community (a partial list of these works is given by , , ). Intensive theoretical activities in this field are presumably related to experimental possibility of observing the macroscopic quantum tunneling in various areas of condensed matter physics. The problem is however not simple and only a limited class of problems have been addressed. Thus even in idealized models one often assumes that the entire system is in thermal equilibrium and attempts to derive quantities of limited value such as the decay rate of the metastable state by an extention of the bounce analysis. In some of these works , , the key quantity is the imaginary part of the free energy, which may be interpreted as the decay rate. However, since equilibrium value of the free energy of the entire system is necessarily real, one must extract the imaginary part by projecting to the initial metastable state. In some discussions in the literature it is not clear how this projection is performed on rigorous grounds, although some of its physical consequences are reasonable. There is indeed some criticism againt this type of approach , . It seems that more fundamental microscopic approach to deal with the tunneling in dissipative medium is needed.
With this background in mind, our aim in the present work is to clarify dynamical aspects of the tunneling in medium: how the barrier penetration basic to quantum tunneling proceeds with real time. The key idea is separation of a small system from a larger environment, and we would like to determine the reduced density matrix for the small system by integrating out environment variables . This makes it easy to compute the penetrating flux factor for an energy eigenstate of the small subsystem. Our approach does not use the Euclidean technique, which in our opinion obscures dynamics of the time evolution. We neither assume that the tunneling system is in thermal equilibrium with environment, although we can discuss this case using our fundamental formula. Moreover, we are able to deal with both the quantum and the thermally activated regions in a unified way.
The model we use to extract exact results for the barrier penetration is the inverted harmonic oscillator (IVHO). Since the form of the potential we use for exact results is very special, we cannot discuss the tunneling for general cases with full confidence. Nevertheless we believe that the method employed in the present work, especially the integral transform of the Wigner function, should be useful to derive approximate, yet valuable results for general potential in the semiclassical approach. We hope to return to a general tunneling potential in our subsequent work.
We take for the environment an infinitely many harmonic oscillators of arbitrary spectrum. This is a standard one studied by many people in the field. The system of a normal harmonic oscillaor coupled to this environment is analytically solvable, as discussed in , . Our barrier penetration model corresponds to a case of imaginary frequency. In the present work we shall employ and extend some techniques we developed in the case of normal oscillator.
The main result of the present work is summarized by a formula for the barrier penetration;
$`f(t)|T(E)|^2,f(t)={\displaystyle \frac{p_0(t)}{\omega _Bq_0(t)}}={\displaystyle \frac{1}{\omega _B}}{\displaystyle \frac{d}{dt}}\mathrm{ln}|q_0(t)|.`$ (1)
This formula is applied to an eigenstate of energy $`E`$ taken for the initial IVHO subsystem. The well known quantum mechanical penetration formula is modified by the environment effect in two ways; first, via the change of the original curvature $`\omega _0`$ (bare parameter) to the renormalized, pole curvature $`\omega _B`$;
$$|T(E)|^2=\frac{1}{1+e^{2\pi |E|/\omega _0}}\frac{1}{1+e^{2\pi |E|/\omega _B}},$$
(2)
with $`E`$ the energy measured from the top of the potential barrier. This effect is essentially similar to, but numerically different from, the curvature renormalization effect much emphasized by Caldeira and Leggett . In their work two cases with and without the friction term, but both including the curvature renormalization given by $`\omega _R`$ (which is larger than $`\omega _B`$) are compared. The result of Caldeira and Leggett is understood when one writes $`\omega _B`$ in terms of $`\omega _R`$.
The second environment effect is the time dependent factor $`f(t)`$ which is a ratio of the momentum $`p_0(t)=\dot{q}_0(t)`$ to the rescaled coordinate trajectory $`q_0(t)`$. This trajectory function $`q_0(t)`$ obyes the homogeneous Langevin equation, eq.(43), under thermal environment, being characterized by the initial energy corresponding to the top of the potential barrier. When the dynamical function $`f(t)=1`$, the IVHO subsystem has the energy of the barrier top, and its deviation from unity is a measure of energy flow from the environment. When $`f(t)>1`$, namely $`|p_0(t)|>\omega _B|q_0(t)|`$, the IVHO system is excited by an energy inflow from the environment. On the other hand, when $`f(t)<1`$, the system is deexcited by an energy outflow. The main new factor $`f(t)`$ deviates from unity within time range of order $`1/\omega _B`$, and both for $`t1/\omega _B`$ and for $`t1/\omega _B`$ the effect is small; $`f(t)1`$. We find interesting examples in which this dynamical function exceeds unity, thus implying enhanced penetration, albeit for a short time of period of order $`1/\omega _B`$.
The rest of this paper is organized as follows. In Section II we explain how we model the environment and its interaction with a quantum system, and introduce the influence functional. Quantum Langevin equation is also briefly touched upon. In Section III we work out exact consequences for the inverted harmonic oscillator, and give the barrier penetration factor taking an energy eigenstate for the initial state. Our general result includes an integral transform of the Wigner function from the initial to the final one, as explained in Appendix A, . The fundamental formula (1) is derived along with an explicit form of $`q_0(t)/q_0(0)`$, eq.(77). The Ohmic or the local approximation (the inherently non-local term of environment interaction in the full Langevin equation being replaced by a few expansion terms) is shown to lead to some anomalous behavior of the dynamical factor. In Section IV some applications to the mixed initial state are discussed using the exact result of the inverted harmonic oscillator. Our understanding of the result of ref. is also made in terms of the diagonalized curvature parameter $`\omega _B`$. Three appendices explain technical points somewhat off the main stream of arguments in the text. Appendix A gives the interesting form of the integral transform of the Wigner function, while Appendix B gives the differential form of the Fokker-Planck equation, derived both for the harmonic model of environment. Appendix C explains parameters necessary for our numerical computation of the dynamical function.
II Model of environment and influence functional method
We expect that the behavior of a small system immersed in thermal environment is insensitive to detailed modeling of the environment and the form of its interaction to the system. Only global quantities such as the environment temperature, the friction, the threshold of the environment spectrum, are expected to be important. Since the pioneering work of Feynman-Vernon and Caldeira-Leggett the standard model uses an infinite set of harmonic oscillators (its coordinate variable denoted by $`Q(\omega )`$) for the environment and a bilinear interaction with the small system (denoted by $`q`$);
$`L_Q={\displaystyle \frac{1}{2}}{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \left(\dot{Q}^2(\omega )\omega ^2Q^2(\omega )\right),L_{\mathrm{int}}=q{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega c(\omega )Q(\omega ).`$ (3)
We assume existence of a threshold $`\omega _c>0`$. The coupling strength to the environment is specified by $`c(\omega )`$. In the present section we do not assume any special form for the potential of the $`q`$system $`V(q)`$.
It is now appropriate to explain the influence functional method . The influence functional denoted by $`[q(\tau ),q^{}(\tau )]`$ results after integration of the environment variable $`Q(\omega )`$ when one computes the density matrix of the entire system. Since the density matrix is a product of the transition amplitude and its conjugate, the path integral formula resulting from the environment integration has a functional dependence both on the system path $`q(\tau )`$ and its conjugate path $`q^{}(\tau )`$. We thus define, when an initial environment is in a mixed state given by a density matrix $`\rho _i(Q_i,Q_i^{}),`$ the influence functional
$`[q(\tau ),q^{}(\tau )]{\displaystyle ๐Q(\tau )๐Q^{}(\tau )๐Q_i๐Q_i^{}}`$
$`{\displaystyle ๐Q_f๐Q_f^{}\delta (Q_fQ_f^{})K(q(\tau ),Q(\tau ))K^{}(q^{}(\tau ),Q^{}(\tau ))\rho _i(Q_i,Q_i^{})},`$ (4)
$`K(q(\tau ),Q(\tau ))=\mathrm{exp}\left(iS_Q[Q]+iS_{\mathrm{int}}[q,Q]\right),`$ (5)
$`S_Q[Q]+S_{\mathrm{int}}[q,Q]={\displaystyle _0^t}d\tau (L_Q[Q]+L_{\mathrm{int}}[q,,Q]),`$ (6)
$`\rho _i(Q_i,Q_i^{})={\displaystyle \underset{n}{}}w_n\psi _n^{}(Q_i^{})\psi _n(Q_i),`$ (7)
where $`w_n`$ is the probability of finding a pure state $`n`$ in the initial environment. The fact that there is a delta function $`\delta (Q_fQ_f^{})`$ for the environment variable at the final time $`t`$ indicates that one does not observe the environment part at $`t`$. Throughout this discussion we consider a definite time interval, $`0<\tau <t`$.
For the thermal ensemble of a $`Q`$ harmonic oscillator of frequency $`\omega `$ ($`\beta =1/T`$ being the inverse temperature), the density matrix is
$`\rho _\beta (Q,Q^{})=`$
$`\left({\displaystyle \frac{\omega }{\pi \mathrm{coth}(\beta \omega /2)}}\right)^{1/2}\mathrm{exp}[{\displaystyle \frac{\omega }{2\mathrm{sinh}(\beta \omega )}}((Q^2+Q^{}{}_{}{}^{2})\mathrm{cosh}(\beta \omega )2QQ^{})],`$ (8)
and one can explicitly perform the $`Q(\omega )`$ path integration in eq.(4), since the $`Q(\omega )`$ integration is Gaussian. The result is a nonlocal action;
$`[q(\tau ),q^{}(\tau )]=\mathrm{exp}[{\displaystyle _0^t}๐\tau {\displaystyle _0^\tau }๐s\left(\xi (\tau )\alpha _R(\tau s)\xi (s)+i\xi (\tau )\alpha _I(\tau s)X(s)\right)],`$
(9)
$`\mathrm{with}\xi (\tau )=q(\tau )q^{}(\tau ),X(\tau )=q(\tau )+q^{}(\tau ),`$ (10)
$`\alpha (\tau )\alpha _R(\tau )+i\alpha _I(\tau )={\displaystyle \frac{1}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐\omega \stackrel{~}{\alpha }(\omega )e^{i\omega \tau },`$ (11)
$`\stackrel{~}{\alpha }(\omega )=i{\displaystyle _{\omega _c}^{\mathrm{}}}d\omega ^{}c^2(\omega ^{})({\displaystyle \frac{1}{\omega ^2\omega ^{}{}_{}{}^{2}+i0^+}}{\displaystyle \frac{2\pi i}{e^{\beta \omega ^{}}1}}\delta (\omega ^{}{}_{}{}^{2}\omega ^2)).`$ (12)
Both of $`\alpha _R`$ and $`\alpha _I`$ are real functions. A complex combination of these, the kernel function $`\alpha (\tau )`$ that appears in the exponent of the influence functional is the real-time thermal Greenโs function for a collection of harmonic oscillators $`Q(\omega )`$, which makes clear a relation to thermal field theory.
The reduced density matrix $`\rho ^{(R)}`$ for the $`q`$ system is defined as follows. For simplicity we take for the initial system a pure quantum state given by a wave function $`\psi (q_i)`$;
$`\rho ^{(R)}(q_f,q_f^{})=`$
$`{\displaystyle ๐q(\tau )๐q^{}(\tau )๐q_i๐q_i^{}\psi ^{}(q_i^{})[q(\tau ),q^{}(\tau )]e^{iS_q[q]iS_q[q^{}]}\psi (q_i)}.`$ (13)
Here $`S_q[q]`$ is the action for the $`q`$system. This density matrix $`\rho ^{(R)}`$ may be used to compute physical quantities of oneโs interest.
It is sometimes convenient to introduce the Wigner function by
$`f_W^{(R)}(x,p)={\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d\xi }{\sqrt{2\pi }}}\rho ^{(R)}(x+{\displaystyle \frac{\xi }{2}},x{\displaystyle \frac{\xi }{2}})e^{ip\xi }.`$ (14)
We shall later mention the master equation for $`f_W^{(R)}`$. Here we quote for comparison a master equation for the Wigner function when the entire system is in a pure quantum state;
$`{\displaystyle \frac{f_W}{t}}=p{\displaystyle \frac{f_W}{x}}+{\displaystyle \frac{1}{i\mathrm{}}}\left(V(x+{\displaystyle \frac{i\mathrm{}}{2}}{\displaystyle \frac{}{p}})V(x{\displaystyle \frac{i\mathrm{}}{2}}{\displaystyle \frac{}{p}})\right)f_W.`$ (15)
It takes a form of infinite dimensional differential equation.
The master equation is simplified in the semiclassical limit of $`\mathrm{}0`$ to
$`{\displaystyle \frac{f_W}{t}}=p{\displaystyle \frac{f_W}{x}}+{\displaystyle \frac{V}{x}}{\displaystyle \frac{f_W}{p}}.`$ (16)
A great virtue of the Wigner function is that this semiclassical equation coincides with the Liouville equation for the distribution function of a classical statistical system in the phase space $`(x,p)`$. It is thus easy to write down the simiclassical solution in the form of an integral transform,
$$f_W(x,p)=_{\mathrm{}}^{\mathrm{}}๐x_i_{\mathrm{}}^{\mathrm{}}๐p_if_W^{(i)}(x_i,p_i)\delta \left(x\stackrel{~}{x}_{\mathrm{cl}}(t)\right)\delta \left(p\stackrel{~}{p}_{\mathrm{cl}}(t)\right),$$
(17)
where the classical deterministic flow, $`(x_i,p_i)(\stackrel{~}{x}_{\mathrm{cl}},\stackrel{~}{p}_{\mathrm{cl}}),`$ is defined by the classical mapping satisfying
$$\stackrel{~}{p}_{\mathrm{cl}}=\frac{d\stackrel{~}{x}_{\mathrm{cl}}}{dt},\frac{d\stackrel{~}{p}_{\mathrm{cl}}}{dt}=\left(\frac{V}{x}\right)_{x=\stackrel{~}{x}_{\mathrm{cl}}}.$$
(18)
Although it is not our main tool of analysis, it might be of some use to recall the quantum Langevin equation for the model of eq.(3) . By eliminating the environment variable $`Q(\omega ,t)`$ one derives the operator equation for the $`q`$ variable;
$`{\displaystyle \frac{d^2q}{dt^2}}+{\displaystyle \frac{V}{q}}+2{\displaystyle _0^t}๐s\alpha _I(ts)q(s)=F_Q(t),`$ (19)
$`F_Q(t)={\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega c(\omega )\left(Q(\omega ,0)\mathrm{cos}(\omega t)+{\displaystyle \frac{P(\omega ,0)}{\omega }}\mathrm{sin}(\omega t)\right),`$ (20)
$`F_Q(\tau )F_Q(s)^{\mathrm{sym}}_{\mathrm{env}}={\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega r(\omega )\mathrm{cos}\omega (\tau s)\mathrm{coth}({\displaystyle \frac{\beta \omega }{2}}),`$ (21)
where $`r(\omega )=c^2(\omega )/(2\omega )`$ and $`_{\mathrm{env}}`$ is the thermal average over the environment variables. Thus, the kernel function $`\alpha _I(\tau )`$ describes a nonlocal action from the environment, while $`F_Q(t)`$ is a random force from the environment. The local approximation to $`\alpha _I`$, taking the form of $`\alpha _I(\tau )=\delta \omega ^2\delta (\tau )+\eta \delta ^{}(\tau ),`$ gives
$$\frac{d^2q}{dt^2}+\frac{V}{q}+\delta \omega ^2q+\eta \frac{dq}{dt}=F_Q(t).$$
(22)
The parameter $`\delta \omega ^2`$ in this case is the frequency shift due to the presence of the environment, and $`\eta `$ is the Ohmic friction. We call this approximation the Ohmic approximation. Perhaps more suitably, it might better be called the local approximation.
On the other hand, the real part of the kernel function $`\alpha _R`$ describes fluctuation, and it is related to the dissipation $`\alpha _I`$ by the fluctuation-dissipation theorem; for their Fourier components
$$\stackrel{~}{\alpha }_I(\omega )=\frac{1}{\pi }๐ซ_{\omega _c}^{\mathrm{}}๐\omega ^{}\frac{2\omega ^{}\stackrel{~}{\alpha }_R(\omega ^{})\mathrm{tanh}(\beta \omega ^{}/2)}{\omega ^2\omega ^{}^2},$$
(23)
where $`๐ซ`$ denotes the principal part of integration.
III Exact results for inverted harmonic oscillator
IIIA Formalism
We specialize the system dynamics of barrier penetration to that of the inverted harmonic oscillator (IVHO) given by the Lagrangian density,
$`L_q={\displaystyle \frac{1}{2}}\dot{q}^2V_q(q),V_q(q)={\displaystyle \frac{1}{2}}\omega _0^2q^2,\omega _0^2>0.`$ (24)
There are similarities to the case of a normal oscillator of $`\omega _0^2<0`$, and we can take over some of the results derived for the unstable ($`|\omega _0|^2>\omega _c^2`$) , or the stable ($`0<|\omega _0|^2<\omega _c^2`$) harmonic oscillator.
The Gaussian nature of the system is a great simplification, as demonstrated by eq.(4), and one may write an effective action for the $`q`$ system including the environment effect;
$`iS_{\mathrm{eff}}`$ $`=`$ $`i{\displaystyle _0^t}๐\tau ({\displaystyle \frac{1}{2}}\dot{\xi }\dot{X}+{\displaystyle \frac{\omega _0^2}{2}}\xi X)`$ (25)
$`{\displaystyle _0^t}๐\tau {\displaystyle _0^\tau }๐s\left(\xi (\tau )\alpha _R(\tau s)\xi (s)+i\xi (\tau )\alpha _I(\tau s)X(s)\right).`$
The obvious linearity in the $`X(\tau )`$ variable here gives a trivial $`X(\tau )`$ path integration in the form of a delta-functional; it determines the $`\xi (\tau )`$ path as
$$\frac{d^2\xi }{d\tau ^2}\omega _0^2\xi (\tau )+2_\tau ^t๐s\xi (s)\alpha _I(s\tau )=0.$$
(26)
The final $`\xi `$path integration then leads to, as the exponent factor,
$$_0^t๐\tau _0^\tau ๐s\xi (\tau )\alpha _R(\tau s)\xi (s),$$
(27)
plus a surface term resulting after the $`X(\tau )`$ partial integration. The function $`\xi (\tau )`$ here is the solution of eq.(26) and the result of path integral must be written with a specified boundary condition, $`\xi (0)=\xi _i,\xi (t)=\xi _f.`$
The standard technique of solving this type of integro-differential equation for $`\xi (\tau )`$ is the Laplace transform . We shall only summarize the final result. Two fundamental solutions to this equation are $`g(t\tau )`$ and its time derivative $`\dot{g}(t\tau )`$ given by
$`g(\tau )`$ $`=`$ $`{\displaystyle \frac{N}{\omega _B}}\mathrm{sinh}(\omega _B\tau )+2{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega H(\omega )\mathrm{sin}(\omega \tau ),`$ (29)
$`N=12{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \omega H(\omega )<1.`$
The important properties are that $`g(\tau )`$ is odd and $`\dot{g}(\tau )`$ is even with $`g(0)=0`$ and $`\dot{g}(0)=1`$ which gives the relation fixing $`N`$. In terms of this basic function $`g(\tau )`$ a general solution to the integro-differential equation with the given boundary condition is
$`\xi (\tau )={\displaystyle \frac{g(t\tau )}{g(t)}}\xi _i+\left(\dot{g}(t\tau ){\displaystyle \frac{\dot{g}(t)}{g(t)}}g(t\tau )\right)\xi _f.`$ (30)
The weight function $`H`$ here is a discontinuity of some analytic function ($`F(z)`$ of a complex variable $`z=\omega `$) across the branch-point singularity along the real axis at $`\omega >\omega _c`$, and is given by
$`H(\omega )={\displaystyle \frac{r(\omega )}{(\omega ^2+\omega _0^2\mathrm{\Pi }(\omega ))^2+(\pi r(\omega ))^2}},r(\omega )={\displaystyle \frac{c^2(\omega )}{2\omega }},`$ (31)
$`\mathrm{\Pi }(\omega )=๐ซ{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega ^{}{\displaystyle \frac{2\omega ^{}r(\omega ^{})}{\omega ^2\omega ^{}^2}},`$ (32)
where the integral for $`\mathrm{\Pi }`$ stands for its principal part. The value $`\omega _B`$ in eq.(29) is determined as a solution for the isolated pole on the real axis at $`\omega ^2=\omega _B^2<0;`$
$`\omega _B^2+\omega _0^2+{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega {\displaystyle \frac{2\omega r(\omega )}{\omega _B^2+\omega ^2}}=0.`$ (33)
In general. one has
$$\omega _B^2>\omega _0^2.$$
(34)
Since many workers in this field use a renormalized potential according to ref., we also introduce these;
$`V_q+V_{qQ}=V_q^{(\mathrm{ren})}(q)+V_{qQ}^{(\mathrm{ren})}(q,Q),`$ (35)
$`V_q^{(\mathrm{ren})}(q)=Vq^2{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega {\displaystyle \frac{r(\omega )}{\omega }}{\displaystyle \frac{1}{2}}\omega _R^2q^2,`$ (36)
$`V_{qQ}^{(\mathrm{ren})}(q,Q)=V_{qQ}(q,Q)+q^2{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega {\displaystyle \frac{r(\omega )}{\omega }}V_{qQ}(q,Q)+{\displaystyle \frac{1}{2}}\delta \omega ^2q^2.`$ (37)
We shall redefine the $`q`$system potential using $`V_q^{(\mathrm{ren})}(q)`$. This renormalized potential gives an inverted harmonic oscillator with the curvature parameter $`\omega _R`$. This renormalized curvature differs from the pole location of the spectral function $`H(\omega )`$, namely $`\omega _B`$ by a factor of order of the interaction coupling. In the weak coupling limit these two quantities are given by
$`\omega _R^2=\omega _0^2+2{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega {\displaystyle \frac{r(\omega )}{\omega }},`$ (38)
$`\omega _B^2\omega _0^2+2{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega {\displaystyle \frac{\omega r(\omega )}{\omega ^2+\omega _0^2}}.`$ (39)
The equation for $`\omega _R^2`$ is a precise definition, while the equation for $`\omega _B^2`$ is an approximate relation valid for the weak coupling, the exact relation being given by eq.(33). In general, one can prove that
$$\omega _B^2<\omega _R^2,$$
(40)
beyond the weak coupling approximation. As an example, in the Ohmic model which will be discussed shortly, the relation becomes
$$\omega _B\omega _R\frac{1}{2}\eta ,$$
(41)
where $`\eta `$ is the Ohmic friction. A more precise relation in this case is eq.(86). The relation to the bare quantity $`\omega _0`$ is
$$\omega _R\omega _0+k\eta ,k\frac{\mathrm{\Omega }}{\pi \omega _0}.$$
(42)
In the infinite cutoff limit, $`\mathrm{\Omega }\mathrm{}`$, the quantity $`k`$ is divergent.
In Fig.1 we show schematically the analytic structure of the function $`F(z^2)`$, basic to the determination of the discontinuity function $`H(\omega )`$. Unlike the case of the normal harmonic oscillator for which the pole location may have an imaginary part, the pole at $`z^2=\omega _B^2`$ appears exactly on the real axis. The other singularity is the branch cut starting from the threshold $`z^2=\omega _c^2`$.
The physical meaning of the basic function $`g(\tau )`$ is better understood by solving the operator Langevin equation for this system;
$`{\displaystyle \frac{d^2q}{dt^2}}\omega _0^2q+2{\displaystyle _0^t}๐\tau \alpha _I(t\tau )q(\tau )=F_Q(\tau ).`$ (43)
The quantity $`\omega _0^2`$ here should be understood as a function of $`\omega _B^2`$ eliminating $`\omega _0^2`$ with eq.(33). The homogeneous solution to this Langevin equation is given by using $`g(t)`$ and $`\dot{g}(t)`$; with the initial data of $`q(0),\dot{q}(0)=p(0),`$
$$q(t)=q(0)\dot{g}(t)+p(0)g(t).$$
(44)
The main term $`g(t)N\mathrm{sinh}(\omega _Bt)/\omega _B`$ and $`\dot{g}(t)N\mathrm{cosh}(\omega _Bt)`$ describes an average motion $`q(t)`$ under the renormalized, inverted harmonic oscillator modified by environment, for which the original parameter $`|\omega _0^2|`$ is replaced by the new shifted $`\omega _B^2`$. Correction to this classical motion given by the second term in eq.(29) describes a damped oscillation with an amplitude decreasing by an inverse power of time at large times .
A straightforward calculation of $`X`$ and $`\xi `$path integration finally gives an effective action valid for any initial state of the $`q`$system. We first define new functions by
$`h(\omega ,t)={\displaystyle _0^t}๐\tau g(\tau )e^{i\omega \tau },`$ (45)
$`k(\omega ,t)={\displaystyle _0^t}๐\tau \dot{g}(\tau )e^{i\omega \tau }=i\omega h(\omega ,t)+g(t)e^{i\omega t}.`$ (46)
With the normalization fixed by unitarity, one has for the effective action $`J`$ defined by $`\rho ^{(R)}=\mathrm{tr}(J\rho _i),`$
$`J(X_f,\xi _f;X_i,\xi _i;t)={\displaystyle \frac{1}{2\pi g(t)}}e^{iS_{\mathrm{cl}}},`$ (47)
where $`S_{\mathrm{cl}}`$ is given by
$`iS_{\mathrm{cl}}`$ $`=`$ $`{\displaystyle \frac{U}{2}}\xi _f^2{\displaystyle \frac{V}{2}}\xi _i^2W\xi _i\xi _f+{\displaystyle \frac{i}{2}}X_f\dot{\xi }_f{\displaystyle \frac{i}{2}}X_i\dot{\xi }_i,`$ (48)
$`U`$ $`=`$ $`({\displaystyle \frac{\dot{g}}{g}})^2I_1+I_22{\displaystyle \frac{\dot{g}}{g}}I_3,V={\displaystyle \frac{1}{g^2}}I_1,`$ (49)
$`W`$ $`=`$ $`{\displaystyle \frac{1}{g}}I_3{\displaystyle \frac{\dot{g}}{g^2}}I_1={\displaystyle \frac{1}{2}}g\dot{V},`$ (50)
$`I_1`$ $`=`$ $`{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \mathrm{coth}{\displaystyle \frac{\beta \omega }{2}}r(\omega )|h(\omega ,t)|^2,`$ (51)
$`I_2`$ $`=`$ $`{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \mathrm{coth}{\displaystyle \frac{\beta \omega }{2}}r(\omega )|k(\omega ,t)|^2,`$ (52)
$`I_3`$ $`=`$ $`{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \mathrm{coth}{\displaystyle \frac{\beta \omega }{2}}r(\omega )\mathrm{}[h(\omega ,t)k^{}(\omega ,t)],`$ (53)
$`\dot{\xi }(\tau )`$ $`=`$ $`\xi _i{\displaystyle \frac{\dot{g}(t\tau )}{g(t)}}\xi _f\left(\stackrel{..}{g}(t\tau ){\displaystyle \frac{\dot{g}(t\tau )\dot{g}(t)}{g(t)}}\right).`$ (54)
For the discussion of the barrier penetration factor, we take a pure initial state given by a wave function $`\psi (x).`$ The Wigner function in this case is
$`f_W^{(R)}(x,p)={\displaystyle \frac{dx_id\xi _i}{2\pi \sqrt{C}}\psi ^{}(x_i\frac{1}{2}\xi _i)\psi (x_i+\frac{1}{2}\xi _i)e^A},`$ (55)
$`A={\displaystyle \frac{\mathrm{det}I}{2C}}\left(\xi _i+i(gJ_1+\dot{g}J_3)(x\dot{g}x_i)+i(\dot{g}J_2+gJ_3)(p\stackrel{..}{g}x_i)\right)^2`$
$`+{\displaystyle \frac{1}{2}}\left(J_1(x\dot{g}x_i)^2+J_2(p\stackrel{..}{g}x_i)^2+2J_3(x\dot{g}x_i)(p\stackrel{..}{g}x_i)\right),`$ (56)
$`C={\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \mathrm{coth}{\displaystyle \frac{\beta \omega }{2}}r(\omega )|\dot{g}h(\omega ,t)gk(\omega ,t)|^2,`$ (57)
$`(J)=(I)^1,J_{1,2}={\displaystyle \frac{I_{2,1}}{I_1I_2I_3^2}},J_3={\displaystyle \frac{I_3}{I_1I_2I_3^2}}.`$ (58)
Although it is not used in the calculation of the flux factor in the next subsection, it is of great theoretical interest to express our result as a transformation of the initial Wigner function $`f_W^{(i)}`$ to the final one $`f_W^{(R)}`$. We give this in Appendix A. This mapping $`f_W^{(i)}f_W^{(R)}`$ is an integrated form describing dynamics of the $`q`$system. Its differential form is known as the Fokker-Planck equation, and we shall explain this briefly in Appendix B.
IIIB Barrier penetration factor
The flux at position $`x`$ is computed from the formula
$$I(x,t)=_{\mathrm{}}^{\mathrm{}}\frac{dp}{\sqrt{2\pi }}pf_W^{(R)}(x,p;t),$$
(59)
to give
$`I(x,t)={\displaystyle }{\displaystyle \frac{dx_id\xi _i}{2\pi g}}\psi ^{}(x_i{\displaystyle \frac{\xi _i}{2}})\psi (x_i+{\displaystyle \frac{\xi _i}{2}})({\displaystyle \frac{\dot{g}}{g}}x+(\stackrel{..}{g}{\displaystyle \frac{\dot{g}^2}{g}})x_i+iW\xi _i)`$
$`\mathrm{exp}\left[{\displaystyle \frac{V}{2}}\left(\xi _i+{\displaystyle \frac{i}{gV}}(x\dot{g}x_i)\right)^2{\displaystyle \frac{1}{2g^2V}}(x\dot{g}x_i)^2\right].`$ (60)
We use the WKB formula for energy eigenstates of IVHO. Considering the incident left mover at $`x<0`$ with the unit flux, we take as the wave funtion at $`x>x_{}`$ (the right turning point)
$`\psi (x){\displaystyle \frac{T(E)}{\sqrt{p(x)}}}\mathrm{exp}\left(i{\displaystyle _x_{}^x}๐x^{}p(x^{})\right),p(x)=\sqrt{\mathrm{\hspace{0.17em}2}E+\omega _B^2x^2},`$ (61)
where $`x_{}=\sqrt{2|E|}/\omega _B`$ and $`T(E)`$ is the transmission coefficient as a function of the energy $`E`$ in a pure quantum state. This choice of the wave function gives the trasmission coefficient,
$$|T(E)|^2\frac{1}{1+e^{\mathrm{\hspace{0.17em}2}\pi E/\omega _B}}.$$
(62)
A point which becomes important when we compare our result with those of other papers is how one prepares the initial state. In many past works a thermal equilibrium between the subsystem and the environment is often assumed, and in this context it is natural to take for our choice of the pure state the reference system characterized by the curvature $`\omega _B`$, the exact pole curvature. The choice of the WKB wave function, using the curvature parameter $`\omega _B`$, fits with this picture. But it should be kept in mind that we may in principle take any reference curvature (hereafter denoted by $`\stackrel{~}{\omega }`$) and in these cases we replace $`\omega _B`$ below by $`\stackrel{~}{\omega }`$.
Expanding wave functions in $`\xi _i`$, we derive a general formula for the dynamical factor. This involves an infinite series of the expansion in $`\xi `$ of the initial density matrix element,
$`\psi (x_i+\xi _i/2)\psi ^{}(x_i\xi _i/2)`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}๐_n(x_i)\xi _i^n.`$ (63)
The first few terms of this series are
$`๐_0(x_i)=|\psi (x_i)|^2,๐_1(x_i)={\displaystyle \frac{1}{2}}(\psi ^{}\psi \psi \psi ^{})(x_i),`$ (64)
$`๐_2(x_i)={\displaystyle \frac{1}{8}}\left[(\psi ^{}^2\psi +\psi ^2\psi ^{})(x_i)+|\psi |^2(x_i)\right].`$ (65)
Computation of the flux factor is then given by
$`I(x,t)={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}I_n(x,t),`$ (66)
$`I_n(x,t)={\displaystyle \frac{dx_id\xi _i}{2\pi g}๐_n(x_i)\xi _i^n\left(\frac{\dot{g}}{g}x+(\ddot{g}\frac{\dot{g}^2}{g})x_i+i\left(\frac{I_3}{g}\frac{\dot{g}I_1}{g^2}\right)\xi _i\right)}`$
$`\times \mathrm{exp}\left[{\displaystyle \frac{V}{2}}\left(\xi _i+{\displaystyle \frac{i}{gV}}(x\dot{g}x_i)\right)^2{\displaystyle \frac{1}{2g^2V}}(x\dot{g}x_i)^2\right]`$ (67)
$`=\left({\displaystyle \frac{g}{\sqrt{I_1}}}\right)^n{\displaystyle \frac{d\alpha }{2\pi \dot{g}}๐_n\left(\frac{x}{\dot{g}}+\frac{\sqrt{I_1}}{\dot{g}}\alpha \right)e^{\alpha ^2/2}}`$
$`\times {\displaystyle }d\beta (\beta +i\alpha )^n\{{\displaystyle \frac{\ddot{g}}{\dot{g}}}x+({\displaystyle \frac{\ddot{g}\sqrt{I_1}}{\dot{g}}}{\displaystyle \frac{I_3}{\sqrt{I_1}}})\alpha +i({\displaystyle \frac{I_3}{g}}{\displaystyle \frac{\dot{g}I_1}{g^2}}){\displaystyle \frac{g}{\sqrt{I_1}}}\beta \}e^{\beta ^2/2}.`$
The second equality follows by a trivial scale change of integration variables $`x_i,\xi _i`$.
One may use expansion of $`๐_n(x/\dot{g}+\sqrt{I_1}\alpha /\dot{g})`$ in powers of coupling,
$`๐_n\left({\displaystyle \frac{x}{\dot{g}}}+{\displaystyle \frac{\sqrt{I_1}}{\dot{g}}}\alpha \right)=`$
$`๐_n\left({\displaystyle \frac{x}{\dot{g}}}\right)+๐_n^{}\left({\displaystyle \frac{x}{\dot{g}}}\right){\displaystyle \frac{\sqrt{I_1}}{\dot{g}}}\alpha +\mathrm{}+{\displaystyle \frac{1}{(k1)!}}๐_n^{(k1)}\left({\displaystyle \frac{x}{\dot{g}}}\right)\left({\displaystyle \frac{\sqrt{I_1}}{\dot{g}}}\alpha \right)^{k1}+\mathrm{}.`$ (68)
For calculation of the penetration factor one only needs $`x\mathrm{}`$ limit of the flux function. From the WKB formula;
$`๐_n(x_i)={\displaystyle \frac{|T(E)|^2}{n!}}{\displaystyle \frac{d^n}{d\xi _i^n}}{\displaystyle \frac{1}{\sqrt{p(x_i\xi _i/2)p(x_i+\xi _i/2)}}}\mathrm{exp}\left(i{\displaystyle _{x_i\xi _i/2}^{x_i+\xi _i/2}}๐x^{}p(x^{})\right)|_{\xi _i=0},`$ (69)
one has
$`๐_n(x_i)|T(E)|^2\left\{i{\displaystyle \frac{(i\omega _Bx_i)^{n1}}{n!}}+๐ช(x_i^{n3})\right\}.`$ (70)
Only the term containing $`\left(\frac{\sqrt{I_1}}{\dot{g}}\alpha \right)^{n1}`$ remains here. One may then derive
$`I_n(x,t)|T(E)|^2(1)^n\left({\displaystyle \frac{\omega _Bg}{\dot{g}}}\right)^{n1}\left({\displaystyle \frac{g\ddot{g}}{\dot{g}^2}}1\right),(n1).`$ (71)
This along with
$$I(\mathrm{},t)=|T(E)|^2f(t),$$
(72)
gives a general formula for the dynamical factor,
$`f(t)={\displaystyle \frac{\ddot{g}}{\omega _B\dot{g}}}\left({\displaystyle \frac{g\ddot{g}}{\dot{g}^2}}1\right){\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{\omega _Bg}{\dot{g}}}\right)^{n1}={\displaystyle \frac{\ddot{g}+\omega _B\dot{g}}{\omega _B(\dot{g}+\omega _Bg)}}.`$ (73)
This is the main result of the present work.
The salient feature of this result is factorization; the main suppression factor given by $`|T(E)|^2`$ is affected by presence of the thermal environment only via renormalization effect, as will be more fully discussed shortly. The other prefactor $`f(t)`$ carries dynamical information of the time evolution.
The basic functions $`g(t)`$ and $`\dot{g}(t)`$ that appear in the dynamical function $`f(t)`$ are the solution to the homogeneous part ($`F_Q=0`$) of the Langevin equation (43) with the initial condition, $`g(0)=0,\dot{g}(0)=1.`$ More conveniently, one may rewrite the dynamical function as
$$f(t)=\frac{p_0(t)}{\omega _Bq_0(t)},p_0(t)=\dot{q}_0(t).$$
(74)
Here $`q_0(t)`$ is the solution with the initial condition,
$$\dot{q}(0)=\omega _Bq(0).$$
(75)
This condition corresponds to the zero energy at time $`t=0`$, since
$`H_q={\displaystyle \frac{1}{2}}\dot{q}^2{\displaystyle \frac{1}{2}}\omega _B^2q^2=0.`$ (76)
Namely, the particle sits on the top of the renormalized potential barrier. Note however that the exact pole curvature $`\omega _B`$ appears here instead of the renormalized $`\omega _R`$. The ratio of the momentum to the coordinate for $`f(t)`$ is a measure of the energy flow from the thermal environment. Thus, the case of $`f(t)>(<)1`$ corresponds to an energy inflow (outflow). Both at $`t1/\omega _B`$ and $`t1/\omega _B`$ this function $`f(t)`$ is nearly unity and it deviates appreciably from unity only within the time range of $`1/\omega _B`$.
An explicit formula useful for detailed analysis of the dynamical function $`f(t)`$ is
$`q_0(t)=\dot{g}(t)+\omega _Bg(t)=Ne^{\omega _Bt}+2{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega H(\omega )\left(\omega \mathrm{cos}\omega t+\omega _B\mathrm{sin}\omega t\right).`$ (77)
The first term represents a simple classical motion under the potential, modified by the curvature renormalization, while the second term is a further deviation due to the environment interaction. The environment effect for $`f(t)`$ appears in two ways; the first via the definition of $`\omega _B`$ determined by the potential renormalization due to the environment interaction. The second one is the continuous part of spectral integral in eq.(77), and its associated deviation of the normalization factor $`N`$ from unity.
Before we go on to specific models of the environment, it is appropriate to discuss some general results. First, both in the weak coupling region and in the asymptotic late time the dynamical function $`f(t)`$ behaves as
$`f(t)f_{\mathrm{asym}}(t),f_{\mathrm{asym}}(t)=1{\displaystyle \frac{2}{\omega _BN}}e^{\omega _Bt}{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega H(\omega )(\omega ^2+\omega _B^2)\mathrm{sin}\omega t.`$ (78)
This form has the correct asymptotic behavior, $`f(\mathrm{})=1`$ as well as the correct initial behavior $`f(0)=1`$ if the integral above is convergent.
On the other hand, one can prove for the initial-time behavior of $`f(t)`$;
$`f(t)=1\omega _B(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{\stackrel{\mathrm{}}{g}(0)}{\omega _B^2}})t+\mathrm{},`$ (79)
$`1{\displaystyle \frac{\stackrel{\mathrm{}}{g}(0)}{\omega _B^2}}={\displaystyle \frac{2}{\omega _B^2}}{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega H(\omega )\omega (\omega ^2+\omega _B^2)>0.`$ (80)
The last inequality means that for very small $`t,`$ $`f(t)<1`$, giving rise to suppression for barrier penetration. In order to have a convergent integral (80) for the formula of $`\stackrel{\mathrm{}}{g}(0)`$ a physical cutoff $`\mathrm{\Omega }`$ of the environment spectrum is necessary, thus the simple-minded Ohmic model without the cutoff should be treated with caution.
The asymptotic formula (78), as will be shown later, describes well, even numerically, the dynamical function at all times, including the small time. For instance, the expansion in time $`t`$ of eq.(78) gives almost the same result as eq.(79), except the missing factor $`N`$.
We would like to stress that two properties, $`f(t)<1`$ with $`f(0)=1`$ for small $`t`$ and $`f(\mathrm{})=1,`$ are generic features for any model having a finite physical cutoff of the spectral weight $`r(\omega )`$. As shown later, however, some approximations or models with the infinite cutoff violate these general properties.
Although our fundamental formula (72) is derived for energy eigenstates, one may compute the barrier penetration factor for any mixed state by suitably weighting this rate for energy eigenstate. A salient feature of our flux formula (72) is that the dependence on the initial (system) state is given by the well known quantum mechanical barrier penetration factor $`|T(E)^2|`$, the other factor $`f(t)`$ being independent of the initial state. This property of factorization is specific to the harmonic barrier. For a more general potential barrier the dynamical factor may depend on the energy of the initial state like $`f(t;E)`$.
We would like to stress that our result extends the result of in several ways. First, we derived the tunneling rate for any energy eigenstate, while the authors of ref. deals with the zero temperature limit of the mixed state in complete equilibrium. Our method is also completely different from the Caldeira-Leggettโs Euclidean approach, and our method makes it possible to discuss dynamics of the time evolution. The third point is that we derived the prefactor rigorously unlike previous approximate calculations. In the next section we explicitly show how we effectively obtain the result of Caldeira and Leggett.
IIIC Some examples
We would like to compute this dynanical function for a few typical examples of the environment spectrum. The spectral function $`r(\omega )`$ is taken as
$`(1)r_O(\omega )={\displaystyle \frac{\eta }{\pi }}\omega ,`$ (81)
$`(2)r_D(\omega )={\displaystyle \frac{\eta \omega }{\pi (1+\omega ^2/\mathrm{\Omega }^2)}},`$ (82)
$`(3)r_T(\omega )={\displaystyle \frac{\eta }{\pi }}(\omega \omega _cฯต(\omega ))\theta (|\omega |\omega _c)\theta (\mathrm{\Omega }|\omega |),`$ (83)
$`(4)r_S(\omega )={\displaystyle \frac{\eta }{\pi }}\omega ^2ฯต(\omega )\theta (\mathrm{\Omega }|\omega |).`$ (84)
As usual, $`\theta (x)`$ is the step function and $`ฯต(x)`$ the signature function. The first one $`r_O(\omega )`$ is the Ohmic model (without physical cutoff of the environment spectrum), the second $`r_D(\omega )`$ the Drude model, and the fourth $`r_S(\omega )`$ a super-Ohmic model, while the third one $`r_T(\omega )`$ has a threshold at $`\omega _c`$. In Fig.2 these spectral weights are schematically depicted. In the last three cases $`\mathrm{\Omega }`$ acts as a cutoff of the environment spectrum. In Appendix C we give parameters necessary to compute the dynamical function.
We note here that the Ohmic model defined here by the spectral weight $`r_O(\omega )`$ should, strictly speaking, be taken as some limit of the infinite cutoff, This cutoff could be a straightforward frequecncy cutoff like $`|\omega |<\mathrm{\Omega }`$, or a smoother one as in the case of the Drude model of a large $`\mathrm{\Omega }`$. The way how the cutoff is introduced does not matter provided a cutoff is there, but in some evaluation of the integral one should first introduce the cutoff and then take the infinite cutoff limit.
The dynamical factor $`f(t)`$ is plotted in Fig. 3$``$7 for these four cases. As the friction $`\eta `$ becomes large, deviation of $`f(t)`$ from unity becomes appreciable, but only in the time range of order $`1/\omega _B`$. For physical reasons we always take $`\eta \omega _B`$. We have found an interesting behavior of $`f(t)`$; some models can give enhancement over the usual quantum formula in the time range $`t1/\omega _B`$. These are the threshold model with a large $`\omega _c`$ and the super-Ohmic model, for which the dynamical factor can exceed unity. The super-Ohmic model has also a peculiar feature that the dynamical function $`f(t)`$ can even become negative for a short time interval. The use of the asymptotic or the weak coupling expression for $`f_{\mathrm{asym}}(t)`$, eq.(78), is compared to the exact result in these figures. Except at small initial times this approximate form gives a reasonable fit to more exact results. Presence of the normalization factor $`N`$ in the formula is important to get a good agreement.
The initial time behavior of the Ohmic model (81) does not reflect the necessary condition of $`f(0)=1`$, since $`f(t)`$ at very small $`t`$ is singular having no smooth derivative at $`t=0`$ (the left and the right derivatives do not meet) in the case of the infinite cutoff limit. This implies that $`f(0)<1`$ in this case indicates anomaly associated with the infinite cutoff, and should not be taken too seriously. On the other hand, the Drude model for a large, but finite cutoff has an expected decrease of $`f(t)`$ from unity at small times. The larger the cutoff is, the more abrupt this decrease is, and a local minimum of $`f(t)`$ appears at a time in proportion to $`1/\mathrm{\Omega }`$.
Finally, we note how the dynamical function behaves in the Ohmic, or the local approximation. We wish to distinguish this Ohmic approximation from the Ohmic model we just discussed. First, write eq.(22) for IVHO;
$$\frac{d^2q}{dt^2}+\eta \frac{dq}{dt}\omega _R^2q=F_Q(t).$$
(85)
There is a problem of how to interpret the zero energy solution since $`\omega _B`$ we need for this is not well defined. One choice might be to use the relation obtained from $`r_O(\omega )`$ in the Ohmic model;
$$\omega _B=\sqrt{\omega _R^2+\frac{\eta ^2}{4}}\frac{\eta }{2}.$$
(86)
This relation is derived by using two exact definitions, the one for $`\omega _B`$ (33) and another for $`\omega _R`$ (38), along with the form of the weight $`r_0(\omega )`$. In this case the dynamical function is trivial; $`f(t)=1`$. But it is by no means obvious that this choice is unique, since without a specific form of the weight function there is no way to locate the pole curvature $`\omega _B`$.
Another choice is to take a more phenomenological view limiting to the local Langevin equation, and define the zero energy condition by taking $`\omega _R`$ for $`\omega _B`$; namely, at time 0,
$$H_q=\frac{1}{2}\dot{q}^2\frac{1}{2}\omega _R^2q^2=0.$$
(87)
The zero energy solution that initially sits on the potential top is then
$$q_0(t)=N\left((\omega _++\omega _R)e^{\omega _{}t}+(\omega _{}\omega _R)e^{\omega _+t}\right).$$
(88)
The parameters here are given by
$`\omega _\pm =\sqrt{\omega _R^2+{\displaystyle \frac{\eta ^2}{4}}}\pm {\displaystyle \frac{\eta }{2}},`$ (89)
both of which are larger than $`\omega _R`$. It is then easy to get
$`f(t)={\displaystyle \frac{\omega _{}(\omega _++\omega _R)e^{\omega _{}t}\omega _+(\omega _{}\omega _R)e^{\omega _+t}}{\omega _R\left((\omega _++\omega _R)e^{\omega _{}t}+(\omega _{}\omega _R)e^{\omega _+t}\right)}}.`$ (90)
This function behaves reasonably at initial times, having $`f(0)=1`$. More precisely,
$$f(t)1\eta t.$$
(91)
Its asymptotic late time behavior is given by
$`f(t)f(\mathrm{})(\mathrm{\hspace{0.17em}1}+Ce^{(\omega _++\omega _{})t}),`$ (92)
$`f(\mathrm{})={\displaystyle \frac{\omega _{}}{\omega _R}}<1,C={\displaystyle \frac{(\omega _++\omega _{})(\omega _R\omega _{})}{\omega _{}(\omega _++\omega _R)}}>0.`$ (93)
The function $`f(t)`$ of (90) is plotted in Fig.8. The property $`f(\mathrm{})<1`$ shows an anomalous behavior of this local approximation, which we regard as an defect inherent in the local approximation.
IV Applications
To illustrate advantages of our approach, we take up two applications of our basic formula, eq.(72) along with (73). The first example is computation of the tunneling rate for the kind of potential depicted in Fig. 9. Qualitatively, a new normal harmonic oscillator is added in the left region of the previous inverted harmonic oscillator. The simplest example of this class of potential is a cubic form,
$$V(x)=\frac{\omega _B^3}{3\sqrt{6V_0}}x^3\frac{\omega _B^2}{2}x^2.$$
(94)
Here $`V_0`$ is the barrier height measured from the left bottom of the potential, and $`\omega _B`$ is taken real and positive. The important quantity added, a new frequency $`\omega _{}`$ in the left harmonic well, is equal to $`\omega _B`$ in this cubic potential. Below we generally distinguish the two, $`\omega _B`$ and $`\omega _{}`$, having a more general tunneling potential in mind.
The problem we set up is to compute the tunneling rate trapped in the metastable state in the left oscillator part at temperature $`T`$, the same temperature as the environment. The setting of this problem is the same as in the Euclidean approach of Grabert et al.. We assume that the temperature $`TV_0`$ with a further condition of $`V_0\omega _{}`$. Under this circumstance the $`n`$-th energy eigenstate of the left oscillator is distributed with the probability $`w_n`$ of the Boltzmann-Gibbs ensemble,
$$w_n=\frac{e^{n\beta \omega _{}}}{_me^{m\beta \omega _{}}}=\left(1e^{\beta \omega _{}}\right)e^{n\beta \omega _{}}.$$
(95)
We then convolute with this weight the barrier penetration factor, to get the tunneling probability $`\mathrm{\Gamma }`$,
$`\mathrm{\Gamma }(T,t)=\left(1e^{\beta \omega _{}}\right){\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}e^{n\beta \omega _{}}|T(V_0+\omega _{}(n+{\displaystyle \frac{1}{2}}))|^2f(t).`$ (96)
We first discuss the infinite time limit, in which $`f(t)1.`$ This probability has the zero temperature limit,
$$\mathrm{\Gamma }(0,\mathrm{})|T(V_0+\frac{\omega _{}}{2})|^2=\frac{1}{1+e^{2\pi (V_0\frac{\omega _{}}{2})/\omega _B}},$$
(97)
which is valid for $`T\omega _{}`$. On the other hand, for $`T\omega _{}`$ the discrete sum
$$\mathrm{\Gamma }(T,\mathrm{})=2\mathrm{sinh}\frac{\beta \omega _{}}{2}e^{\beta V_0}\underset{n=0}{\overset{\mathrm{}}{}}\frac{e^{\beta \omega _{}(n+\frac{1}{2}V_0/\omega _{})}}{1+e^{\mathrm{\hspace{0.17em}2}\pi ((n+\frac{1}{2})\omega _{}V_0)/\omega _B}},$$
(98)
holds. This approximately reduces to
$`\mathrm{\Gamma }(T,\mathrm{})2{\displaystyle \frac{\beta \omega _{}}{2}}e^{\beta V_0}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dx}{\beta \omega _{}}}{\displaystyle \frac{1}{1+x^{2\pi /\beta \omega _B}}}`$
(99)
$`={\displaystyle \frac{\beta \omega _Be^{\beta V_0}}{2\mathrm{sin}(\beta \omega _B/2)}}e^{\beta V_0}.`$ (100)
The last formula holds for $`T\omega _B`$. This is the expected classical formula for the barrier penetration at finite temperature. More quantitatively we numerically computed eq.(96) to compare with various approximate formulas. The quantity $`\mathrm{\Gamma }(T,\mathrm{})`$, computed from eq.(96) with $`f(t)=1`$, is plotted in Fig.10. In this figure an approximate formula,
$$\mathrm{\Gamma }(T,\mathrm{})=\frac{\omega _B\mathrm{sinh}(\beta \omega _{}/2)}{\omega _{}\mathrm{sin}(\beta \omega _B/2)}e^{\beta V_0}+\frac{1}{1+e^{\mathrm{\hspace{0.17em}2}\pi (\omega _{}/(2\omega _B)V_0/\omega _B)}},$$
(101)
is compared to the exact result in the case of $`\omega _B=\omega _{}`$. This interpolation formula is a simple sum of the improved high and the improved low temperature limit.
We now present our understanding of the result of Caldeira and Leggett for the cubic potential $`V(x)`$ (94). For this we extend our result of IVHO to this class of potential by a new formula,
$$I(\mathrm{},t)=\mathrm{exp}\left(\mathrm{\hspace{0.17em}2}_{x_1}^{x_2}๐x\sqrt{2(V(x)E)}\right)\frac{1}{\omega _B}\frac{d}{dt}\mathrm{ln}|q_0(t)|,$$
(102)
where $`x_i`$ are turning points for the energy $`E=V_0`$ solution. We take as the reference curvature $`\stackrel{~}{\omega }=\omega _B`$. The trajectory function $`q_0(t)`$ is taken as the classical, zero energy solution, but we ignore the dynamical function since it is almost unity in the equilibrium circumstance of . Thus we find for the tunneling probability to be compared
$$I(\mathrm{},t)|T(E)|^2=\mathrm{exp}\left(\frac{36}{5}\frac{V_0}{\omega _B}\right).$$
(103)
The authors of write the tunneling probability in terms of the renormalized curvature $`\omega _R`$ and the friction $`\eta `$. For the Ohmic model of small friction, we get (using eq.(86) with $`\eta \omega _R`$),
$$|T(E)|^2e^{\mathrm{\hspace{0.17em}2}\pi V_0/\omega _R}e^{\mathrm{\Delta }B}$$
(104)
with
$$\mathrm{\Delta }B\frac{18}{5}\frac{\eta V_0}{\omega _R^2}=3.6\frac{\eta V_0}{\omega _R^2}.$$
(105)
If we had used the IVHO potential instead of the cubic form, we would have obtained a numerical factor slightly different from this value, $`3.6\pi 3.14`$. On the other hand, Caldeira and Leggett obtain, using a different method,
$$\mathrm{\Delta }B_{\mathrm{CL}}\frac{162\zeta (3)}{\pi ^3}\frac{\eta V_0}{\omega _R^2}6.28045\frac{\eta V_0}{\omega _R^2}.$$
(106)
Thus our result gives a result numerically different by a factor of about 2 in the weak coupling limit.
On the other hand, in the large coupling limit, $`\eta 2\omega _R`$, these two are
$$\mathrm{\Delta }B=\frac{36}{5}\frac{\eta V_0}{\omega _R^2},\mathrm{\Delta }B_{\mathrm{CL}}=3\pi \frac{\eta V_0}{\omega _R^2}.$$
(107)
(Although our notation here suggests that this is a correction term to the main term of $`e^{\mathrm{\hspace{0.17em}2}\pi V_0/\omega _R}`$, it is actually a leading term for the case of $`\eta 2\omega _R`$.) Our result is about 0.8 times the Caldeira-Leggett value.
In general, when one writes the deviation of the penetration factor from the case of no environment as $`e^{\mathrm{\Delta }B}`$, one has the form,
$$\mathrm{\Delta }B=\mathrm{\Phi }\left(\frac{\eta }{2\omega _R}\right)\frac{\eta V_0}{\omega _R^2}.$$
(108)
Both in our case and in Caldeira and Leggett the function $`\mathrm{\Phi }(\alpha )`$ is slowly varying albeit numerically different, and the deviation factor is dominated by $`\eta V_0/\omega _R^2`$. This appears to be the most important dependence of parameters, the numerical factor being a secondary effect.
In our interpretation of the result of Caldeira and Leggett it is crucial to use the pole curvature $`\omega _B`$ as our reference, and this choice is reasonable, because it corresponds to an equilibrium in the zero temperature limit considered in . With the dynamical function taken as $`f(t)=1`$, this is the only way how the friction ($`\eta `$) dependence can appear in our approach, namely via the parameter $`\omega _B`$ in the initial density matrix. It is important that our inequality $`\omega _B^2<\omega _R^2`$ implies the general result of suppressed rate of tunneling in medium, the main point stressed by Caldeira and Leggett. We also note that detailed comparison between our result and that of is possible only by assuming the relation (86), specific to the Ohmic spectrum. For different models the difference between the two might be larger.
In future we wish to examine further this discrepancy after we develop formalism for more general potentials.
The second application is the problem of time evolution for the same potential as above. For this we consider the temperature range of $`\omega _{}TV_0`$ such that the average energy eigenstate (of $`\overline{E}T`$) may be treated semiclassically. We thus regard a particle in the $`n`$-th energy level ($`n1`$) of the left harmonic oscillator as moving almost classically with the periodic motion of frequency $`\omega _{}`$. Each time this particle hits against the right barrier, it has a prescribed probability (72) of tunneling into the far right region. Starting at time $`t=0`$, one counts the $`k`$-th encounter with the barrier at time $`t=2\pi k/\omega _{}`$ until the final time $`t_f`$ such that there are roughly $`t_f\omega _{}/2\pi `$ times of possibilities of the barrier penetration. We find it reasonable to use a reset time for each encounter for the time $`t`$ of eq.(72). Summing up all these possibilities, one gets the total tunneling probability,
$`\mathrm{\Gamma }_n(t_f)=|T(E_n)|^2{\displaystyle \frac{t_f\omega _{}}{2\pi }}\overline{f},`$ (109)
where $`\overline{f}`$ is some sort of representative value for the dyanamical factor for each encounter, perhaps some average of $`f(t)`$ (73) over one period of oscillation under the harmonic well like
$$\overline{f}=\frac{1}{t_{}}_0^t_{}๐tf(t),t_{}=\frac{\omega _{}}{2\pi }.$$
(110)
Another choice for $`\overline{f}`$ is the dynamical function at some particular value of time during one period of oscillation, for instance at the classical turning point.
For $`\omega _B\omega _{},`$ $`\overline{f}1`$, and
$$\mathrm{\Gamma }_n(t)|T(E_n)|^2\frac{t\omega _{}}{2\pi }.$$
(111)
Thus, the total probability $``$ time $`t`$, and one may define the tunneling rate per unit time,
$$\frac{\mathrm{\Gamma }_n(t)}{t}=\frac{\omega _{}}{2\pi }|T(E)|^2.$$
(112)
This is nothing but the classic Kramers formula .
On the other hand, for $`\omega _B\omega _{}`$ there may be a large deviation from the quantum mechanical formula. We plot in Fig.11 some examples of the factor $`\overline{f}`$ (110) as a function of the average time $`t_{}`$. In most cases studied, $`\overline{f}`$ is some fraction of unity, typically larger than 0.5.
Appendix A Integral transform of the Wigner function
After some algebra, we obtain
$`f_W^{(R)}(x,p)={\displaystyle \frac{1}{2\pi \sqrt{I_1I_2I_3^2}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐x_i{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐p_if_W^{(i)}(x_i,p_i)`$
$`\mathrm{exp}\left[{\displaystyle \frac{1}{2(I_1I_2I_3^2)}}\left(I_1(pp_{\mathrm{cl}})^2+I_2(xx_{\mathrm{cl}})^22I_3(xx_{\mathrm{cl}})(pp_{\mathrm{cl}})\right)\right],`$ (113)
$`x_{\mathrm{cl}}=\dot{g}x_i+gp_i,p_{\mathrm{cl}}=\stackrel{..}{g}x_i+\dot{g}p_i.`$ (114)
The definitions of $`I_i`$ are given in the text, eqs.(51)$``$(53). The time dependent functions, $`x_{\mathrm{cl}}(t),p_{\mathrm{cl}}(t)`$, are homogeneous solutions to the Langevin equation (43) with $`F_Q=0`$.
One may view the mapping from $`f_W^{(i)}`$ to $`f_W^{(R)}`$ as a kind of fluid flow. Compared to the classical mapping given in Section II, the quantum solution (113) in thermal medium is not deterministic with a broadening given by the coefficient matrix,
$$(I)=\left(\begin{array}{cc}I_1& I_3\\ I_3& I_2\end{array}\right).$$
(115)
Moreover, the initial distribution $`f_W^{(i)}`$ itself is broadened by quantum mechanical effects. The peak point of the distribution is at $`(x_{\mathrm{cl}}(t),p_{\mathrm{cl}}(t)).`$ One might imagine that the mapping $`(x_i,p_i)(x_{\mathrm{cl}}(t),p_{\mathrm{cl}}(t))`$ is not invertible due to dissipative effects from the environment. This is not true; the mapping is actually invertible and
$`x_i={\displaystyle \frac{\dot{g}x_{\mathrm{cl}}gp_{\mathrm{cl}}}{\dot{g}^2g\stackrel{..}{g}}},p_i={\displaystyle \frac{\stackrel{..}{g}x_{\mathrm{cl}}+\dot{g}p_{\mathrm{cl}}}{\dot{g}^2g\stackrel{..}{g}}},`$ (116)
with $`\dot{g}^2g\stackrel{..}{g}0`$.
Appendix B Fokker-Planck equation
One may derive the master equation for the reduced density matrix as described in ref.. Our formula for the Wigner function (113) is considered as an explicit and convenient solution to this type of master equation. From the master equation one can derive a Fokker-Planck equation for the Wigner function;
$`{\displaystyle \frac{f_W^{(R)}}{t}}`$ $`=`$ $`p{\displaystyle \frac{f_W^{(R)}}{x}}+\mathrm{\Omega }^2(t)x{\displaystyle \frac{f_W^{(R)}}{p}}+C(t){\displaystyle \frac{}{p}}(pf_W^{(R)})2D_{pp}(t){\displaystyle \frac{^2f_W^{(R)}}{p^2}}`$ (117)
$`+`$ $`2D_{xp}(t){\displaystyle \frac{^2f_W^{(R)}}{xp}},`$
$`\mathrm{\Omega }^2(t)`$ $`=`$ $`{\displaystyle \frac{\stackrel{2}{\stackrel{..}{g}}\dot{g}\stackrel{\mathrm{}}{g}}{g\stackrel{..}{g}\dot{g}^2}}={\displaystyle \frac{\dot{g}}{g}}{\displaystyle \frac{d}{dt}}\mathrm{ln}\left({\displaystyle \frac{g}{\dot{g}}}(\stackrel{..}{g}{\displaystyle \frac{\dot{g}^2}{g}})\right),`$ (118)
$`C(t)`$ $`=`$ $`{\displaystyle \frac{g\stackrel{\mathrm{}}{g}\dot{g}\stackrel{..}{g}}{g\stackrel{..}{g}\dot{g}^2}},`$ (119)
$`D_{pp}(t)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{g\stackrel{\mathrm{}}{g}\dot{g}\stackrel{..}{g}}{g\stackrel{..}{g}\dot{g}^2}}U+{\displaystyle \frac{\dot{g}}{2g}}{\displaystyle \frac{g^2\stackrel{\mathrm{}}{g}2g\dot{g}\stackrel{..}{g}+\dot{g}^3}{g\stackrel{..}{g}\dot{g}^2}}W{\displaystyle \frac{\dot{U}}{2}}\dot{g}\dot{W}\right),`$ (120)
$`D_{xp}(t)`$ $`=`$ $`Ug\dot{W}+{\displaystyle \frac{g^2\stackrel{\mathrm{}}{g}2g\dot{g}\stackrel{..}{g}+\dot{g}^3}{g\stackrel{..}{g}\dot{g}^2}}W,`$ (121)
where coefficient functions, $`\mathrm{\Omega }^2(t),C(t),D_{pp}(t),D_{xp}(t)`$ are local functions of time and are written in terms of $`g(t),U,V,W`$.
Quantities that appear in this equation are well understood by writing a set of moment equations of low orders;
$`{\displaystyle \frac{dx}{dt}}`$ $`=`$ $`p,{\displaystyle \frac{dp}{dt}}=\mathrm{\Omega }^2(t)xC(t)p,`$ (122)
$`{\displaystyle \frac{dx^2}{dt}}`$ $`=`$ $`2xp,`$ (123)
$`{\displaystyle \frac{dp^2}{dt}}`$ $`=`$ $`\mathrm{\hspace{0.17em}2}\mathrm{\Omega }^2(t)xp2C(t)p^24D_{pp}(t),`$ (124)
$`{\displaystyle \frac{dxp}{dt}}`$ $`=`$ $`p^2\mathrm{\Omega }^2(t)x^2C(t)xp+2D_{xp}(t).`$ (125)
For instance, the quantity $`\mathrm{\Omega }^2(t)`$ here is a time dependent curvature parameter modified from the original $`\omega _0^2`$ to that in thermal medium, while $`C(t)`$ is a time dependent friction. In similar fashons one understands $`D_{pp}`$ and $`D_{xp}`$ as fluctuations. Physical behaviors of the harmonic oscillator system under thermal environment are all determined by these four quantities which are functions of the local time $`t`$.
Limiting values relevant to large times, $`t1/\omega _B`$, are
$`\mathrm{\Omega }^2(t)\omega _B^2,C(t)0,`$ (126)
$`D_{pp}(t){\displaystyle \frac{\omega _B}{4}}\dot{g}W{\displaystyle \frac{\omega _B^3}{4}}{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \mathrm{cosh}({\displaystyle \frac{\beta \omega }{2}})H(\omega ),`$ (127)
$`D_{xp}(t)U\dot{g}W{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega \mathrm{cosh}({\displaystyle \frac{\beta \omega }{2}})(\omega ^2+{\displaystyle \frac{5}{4}}\omega _B^2)H(\omega ).`$ (128)
Appendix C Parameters in specific environment models
We give various parameters in four specific models of environment given in Section IIIC. These parameters are used to calculate $`g(t)`$ according to
$`g(t)`$ $`=`$ $`{\displaystyle \frac{N}{\omega _B}}\mathrm{sinh}(\omega _Bt)+2{\displaystyle _{\omega _c}^{\mathrm{}}}๐\omega H(\omega )\mathrm{sin}(\omega t),`$ (129)
$`H(\omega )`$ $`=`$ $`{\displaystyle \frac{r(\omega )}{(\omega ^2+\omega _R^2\mathrm{\Pi }(\omega ))^2+\pi ^2r(\omega )^2}},`$ (130)
$`\omega _R^2`$ $`=`$ $`\omega _B^2+C(\omega _B).`$ (131)
The parameters are given as follows.
(1) Ohmic model
$`N=\left(1+{\displaystyle \frac{\eta }{2\omega _B}}\right)^1,C(\omega _B)=\eta \omega _B,\mathrm{\Pi }(\omega )=0,`$ (132)
$`g(t)={\displaystyle \frac{N}{2\omega _B}}\left(e^{\omega _Bt}e^{(\omega +\eta )t}\right).`$ (133)
(2) Drude model
$`N`$ $`=`$ $`\left(1+{\displaystyle \frac{\eta }{2\omega _B}}{\displaystyle \frac{\mathrm{\Omega }^2}{(\omega _B+\mathrm{\Omega })^2}}\right)^1,`$ (134)
$`C(\omega _B)`$ $`=`$ $`{\displaystyle \frac{\eta \omega _B\mathrm{\Omega }}{\omega _B+\mathrm{\Omega }}},\mathrm{\Pi }(\omega )={\displaystyle \frac{\eta \omega ^2\mathrm{\Omega }}{\omega ^2+\mathrm{\Omega }^2}},`$ (135)
$`g(t)`$ $`=`$ $`{\displaystyle \frac{N}{\omega _B}}\mathrm{sinh}(\omega _Bt)+{\displaystyle \frac{\eta \mathrm{\Omega }^2}{(\alpha _+^2\omega _B^2)(\alpha _{}^2\omega _B^2)(\alpha _+^2\alpha _{}^2)}}`$ (136)
$`\times `$ $`\left((\alpha _+^2\alpha _{}^2)e^{\omega _Bt}+(\alpha _{}^2\omega _B^2)e^{\alpha _+t}+(\omega _B^2\alpha _+^2)e^{\alpha _{}t}\right),`$
$`\alpha _\pm `$ $`=`$ $`{\displaystyle \frac{\omega _B+\mathrm{\Omega }}{2}}\pm \sqrt{{\displaystyle \frac{(\omega _B\mathrm{\Omega })^2}{4}}{\displaystyle \frac{\eta \mathrm{\Omega }^2}{\omega _B+\mathrm{\Omega }}}}.`$ (137)
(3) Threshold model
$`N`$ $`=`$ $`\left(1{\displaystyle \frac{\eta }{\pi }}{\displaystyle \frac{\mathrm{\Omega }\omega _c}{\mathrm{\Omega }^2+\omega _B^2}}+{\displaystyle \frac{\eta }{\pi \omega _B}}\mathrm{arctan}{\displaystyle \frac{\mathrm{\Omega }}{\omega _B}}{\displaystyle \frac{\eta }{\pi \omega _B}}\mathrm{arctan}{\displaystyle \frac{\omega _c}{\omega _B}}\right)^1,`$ (138)
$`C(\omega _B)`$ $`=`$ $`{\displaystyle \frac{2\eta \omega _B}{\pi }}\left(\mathrm{arctan}{\displaystyle \frac{\mathrm{\Omega }}{\omega _B}}\mathrm{arctan}{\displaystyle \frac{\omega _c}{\omega _B}}\right)`$ (139)
$`+`$ $`{\displaystyle \frac{\eta \omega _c}{\pi }}\left\{\mathrm{ln}\left(1+{\displaystyle \frac{\omega _B^2}{\mathrm{\Omega }^2}}\right)\mathrm{ln}\left(1+{\displaystyle \frac{\omega _B^2}{\omega _c^2}}\right)\right\},`$
$`\mathrm{\Pi }(\omega )`$ $`=`$ $`{\displaystyle \frac{\eta (\omega \omega _c)}{2\pi }}\mathrm{ln}({\displaystyle \frac{\omega }{\omega _c}}1)^2{\displaystyle \frac{\eta (\omega +\omega _c)}{2\pi }}\mathrm{ln}({\displaystyle \frac{\omega }{\omega _c}}+1)^2`$ (140)
$``$ $`{\displaystyle \frac{\eta (\omega \omega _c)}{2\pi }}\mathrm{ln}({\displaystyle \frac{\omega }{\mathrm{\Omega }}}1)^2+{\displaystyle \frac{\eta (\omega +\omega _c)}{2\pi }}\mathrm{ln}({\displaystyle \frac{\omega }{\mathrm{\Omega }}}+1)^2.`$
(4) Super-Ohmic model
$`N`$ $`=`$ $`\left(1{\displaystyle \frac{\eta }{\pi }}{\displaystyle \frac{\mathrm{\Omega }^2}{\mathrm{\Omega }^2+\omega _B^2}}+{\displaystyle \frac{\eta }{\pi }}\mathrm{ln}{\displaystyle \frac{\mathrm{\Omega }^2+\omega _B^2}{\omega _B^2}}\right)^1,`$ (141)
$`C(\omega _B)`$ $`=`$ $`{\displaystyle \frac{\eta \omega _B^2}{\pi }}\mathrm{ln}{\displaystyle \frac{\mathrm{\Omega }^2+\omega _B^2}{\omega _B^2}},\mathrm{\Pi }(\omega )={\displaystyle \frac{\eta \omega ^2}{\pi }}\mathrm{ln}{\displaystyle \frac{\mathrm{\Omega }^2\omega ^2}{\omega ^2}}.`$ (142)
In calculation of $`g(t)`$ for the Ohmic model one needs a frequency cutoff in intermediate steps of integration, but the final result does not depend on this cutoff factor. For the model having a threshold and for the suer-Ohmic model we cannot get analytic forms of the basic function $`g(t)`$.
Acknowledgment
The work of Sh. Matsumoto is partially supported by the Japan Society of the Promotion of Science.
Figure caption
Fig.1
Analytic structure giving the spectrum. The pole at $`\omega _B^2`$ moves as indicated when the friction becomes large with $`\omega _R^2`$ fixed.
Fig.2
Schematic form of model spectral weights.
Fig.3
Dynamical function for the Ohmic model. Values of the friction relative to the system curvature are given for each case. Dotted lines are calculated using the approximate, asymptotic formula $`f_{\mathrm{asym}}(t)`$, eq.(78), in the text.
Fig.4
Dynamical function for the Drude model. Values of the friction and the cutoff relative to the system curvature are given for each case. Dotted lines are calculated using the approximate, asymptotic formula $`f_{\mathrm{asym}}(t)`$, eq.(78), in the text.
Fig.5
Dynamical function for the threshold model of low threshold. Values of the friction, the cutoff, and the threshold relative to the system curvature are given for each case. Dotted lines are calculated using the approximate, asymptotic formula $`f_{\mathrm{asym}}(t)`$, eq.(78), in the text.
Fig.6
The same as in Fig.5, for the threshold model of high threshold.
Fig.7
Dynamical function for the super-Ohmic model. Values of the friction and the cutoff are given for each case. Dotted lines are calculated using the approximate, asymptotic formula $`f_{\mathrm{asym}}(t)`$, eq.(78), in the text.
Fig.8
Dynamical function in the local, Ohmic approximation. Values of the friction are given for each case.
Fig.9
Schematic form of a general potential.
Fig.10
Thermally averaged tunneling probability based on eq.(96) with $`f(t)=1`$ in the text taking $`\omega _{}=\omega _B`$. Crossed points are calculated using the approximate formula (101).
Fig.11
Time averaged dynamical factor for a few models.
|
warning/0006/math0006087.html
|
ar5iv
|
text
|
# Virasoro algebra and wreath product convolution
## Introduction
It is by now well known that a direct sum $`_{n0}R(S_n)`$ of the Grothendieck rings of symmetric groups $`S_n`$ can be identified with the Fock space of the Heisenberg algebra of rank one. One can construct vertex operators whose components generate an infinite-dimensional Clifford algebra, the relation known as boson-fermion correspondence \[F\] (also see \[J\]). A natural open problem which arises here is to understand the group theoretic meaning of more general vertex operators in a vertex algebra \[B, FLM\].
A connection between a direct sum $`R_\mathrm{\Gamma }=_{n0}R(\mathrm{\Gamma }_n)`$ of the Grothendieck rings of wreath products $`\mathrm{\Gamma }_n=\mathrm{\Gamma }S_n`$ associated to a finite group $`\mathrm{\Gamma }`$ and vertex operators has been realized recently in \[W\] and \[FJW\] (also see \[Z, M\] for closely related algebraic structures on $`R_\mathrm{\Gamma }`$). When $`\mathrm{\Gamma }`$ is trivial one recovers the above symmetric group picture. On the other hand, this wreath product approach turns out to be very much parallel to the development in the theory of Hilbert schemes of points on a surface (cf. \[W, N2\] and references therein). As one expects that new insight in one theory will shed new light on the other, this refreshes our hope of understanding the group theoretic meaning of general vertex operators.
The goal of this paper is to take the next step in this direction to produce the Virasoro algebra within the framework of wreath products. Denote by $`\mathrm{\Gamma }_{}`$ the set of conjugacy classes of $`\mathrm{\Gamma }`$, and by $`c^0`$ the identity conjugacy class. Recall \[M, Z\] that the conjugacy classes of the wreath product $`\mathrm{\Gamma }_n`$ are parameterized by the partition-valued functions on $`\mathrm{\Gamma }_{}`$ of length $`n`$ (also see Section 2). Given $`c\mathrm{\Gamma }_{}`$, we denote by $`\lambda _c`$ the function which maps $`c`$ to the one-part partition $`(2)`$, $`c^0`$ to the partition $`(1^{n2})`$ and other conjugacy classes to $`0`$. We will define an operator $`\mathrm{\Delta }_c`$ in terms of the convolution with the characteristic class function on $`\mathrm{\Gamma }_n`$ (for all $`n`$) associated to the conjugacy class parameterized by $`\lambda _c`$. We show that this operator can be identified with a differential operator which is the zero mode of a certain vertex operator when we identify $`R_\mathrm{\Gamma }`$ as in \[M\] with a symmetric algebra. A group theoretic construction of Heisenberg algebra has been given in \[W\] (also see \[FJW\]) which acts on $`R_\mathrm{\Gamma }`$ irreducibly. The commutator between $`\mathrm{\Delta }_c`$ and the Heisenberg algebra generators on $`R_\mathrm{\Gamma }`$ provides us the Virasoro algebra generators.
Our construction is motivated in part by the work of Lehn \[L\] in the theory of Hilbert schemes. Among other results, he showed that an operator defined in terms of intersection with the boundary of Hilbert schemes may be used to produce the Virasoro algebra when combined with earlier construction of Heisenberg algebra due to Nakajima and independently Grojnowski \[N1, Gr\]. It remains an important open problem to establish a precise relationship between Lehnโs construction and ours.
We remark that the convolution operator in the symmetric group case (i.e. $`\mathrm{\Gamma }`$ trivial), when interpreted as an operator on the space of symmetric functions, is intimately related to the Hamiltonian in Calegero-Sutherland integrable system and to the Macdonald operator which is used to define Macdonald polynomials \[AMOS, M\].
After we discovered our group theoretic approach toward the Virasoro algebra, we notice that our convolution operator in the symmetric group case has been considered by Goulden \[Go\] when studying the number of ways of writing permutations in a given conjugacy class as products of transpositions. We regard this as a confirmation of our belief that the connections between $`R_\mathrm{\Gamma }`$ and (general) vertex operators are profound. It is likely that often time when we understand something deeper in this direction, we may realize that it is already hidden in the vast literature of combinatorics, particularly on symmetric groups and symmetric functions for totally different needs. Then the virtue of our point of view will be to serve as a unifying principle which patches together many pieces of mathematics which were not suspected to be related at all.
The plan of this paper is as follows. In Sect. 1, we recall some basics of Heisenberg and Virasoro algebras from the viewpoint of vertex algebras. In Sect. 2 we set up the background in the theory of wreath products which our main constructions are based on. In Sect. 3 we present our main results. Some materials in the paper are fairly standard to experts, but we have decided to include them in hope that it might be helpful to the reader with different backgrounds.
## 1 Basics of Heisenberg and Virasoro algebras
In the following we will present some basic constructions in vertex algebras which give us the Virasoro algebra from a Heisenberg algebra.
Let $`L`$ be a rank $`N`$ lattice endowed with an integral non-degenerate symmetric bilinear form. Indeed we will only need the case $`L=Z^N`$ with the standard bilinear form.
Denote by $`h=C_ZL`$ the vector space generated by $`L`$ with the bilinear from $`,`$ induced from $`L`$. We define the Heisenberg algebra
$$\widehat{h}=C[t,t^1]hCC$$
with the following commutation relations:
$`[C,a_n]`$ $`=`$ $`0,`$
$`[a_n,b_m]`$ $`=`$ $`n\delta _{n,m}a,bC,`$
where $`a_n`$ denotes $`t^na,ah`$.
We denote by $`S_L`$ the symmetric algebra generated by $`\widehat{h}^{}=t^1C[t^1]h`$. It is well known that $`S_L`$ can be given the structure of an irreducible module over Heisenberg algebra $`\widehat{h}`$ by letting $`a_n,n>0`$ acts as multiplication and letting
$`a_n.a_{n_1}^1a_{n_2}^2\mathrm{}a_{n_k}^k={\displaystyle \underset{i=1}{\overset{k}{}}}\delta _{n,n_i}a,a^ia_{n_1}^1a_{n_2}^2\mathrm{}\stackrel{ห}{a}_{n_i}^i\mathrm{}a_{n_k}^k,`$
where $`n0,n_i>0,a,a^ih`$ for $`i=1,\mathrm{},k`$, and $`\stackrel{ห}{a}_{n_i}^i`$ means the very term is deleted. A natural gradation on $`S_L`$ is defined by letting
$$\mathrm{deg}(a_{n_1}^1a_{n_2}^2\mathrm{}a_{n_k}^k)=n_1+\mathrm{}+n_k.$$
We say an operator on $`S_L`$ is of degree $`p`$ if it maps any $`n`$-th graded subspace of $`S_L`$ to $`(n+p)`$-th graded subspace.
The space $`S_L`$ carries a natural structure of a vertex algebra \[B, FLM\]. It is convenient to use the generating function in a variable $`z`$:
$$a(z)=\underset{nZ}{}a_nz^{n1},ah.$$
In the language of vertex algebras, this is the vertex operator associated to the vector $`a_1S_L`$. The normally ordered product between two vertex operators $`a(z)`$ and $`b(z)`$ is defined as
$$:a(z)b(z):=\underset{n<0}{}a_nz^{n1}b(z)+b(z)\underset{n0}{}a_nz^{n1}.$$
We remark that the commutation relations between $`a_n,b_m,n,mZ`$ are encoded in the following operator product expansion (OPE) (cf. \[FLM\]):
$$a(z)b(w)\frac{a,b}{(zw)^2}.$$
We recall that the Virasoro algebra is spanned by $`L_n,nZ`$ and a central element $`C`$, with the following commutation relations:
$$[L_n,L_m]=(nm)L_{n+m}+\frac{n^3n}{12}\delta _{n,m}C.$$
It is convenient to denote
$$L(z)=\underset{nZ}{}L_nz^{n2}.$$
Given an orthonormal basis $`a^i,i=1,\mathrm{},n`$ of $`h`$, we define a series of operators $`\stackrel{~}{L}_n^i`$ $`(i=1,\mathrm{}N,nZ)`$ acting on $`S_L`$ by
$`\stackrel{~}{L}^i(z){\displaystyle \underset{nZ}{}}\stackrel{~}{L}_n^iz^{n2}={\displaystyle \frac{1}{2}}:a^i(z)a^i(z):.`$
The following proposition is well known (cf. e.g. \[FLM\]).
###### Proposition 1
The operators $`\stackrel{~}{L}_n^i`$ $`(i=1,\mathrm{}N,nZ)`$ generate $`N`$ commutative copies of Virasoro algebra of central charge $`1`$. The operators $`\stackrel{~}{L}_n,nZ`$ generate the Virasoro algebra with central charge $`N`$. Namely we have
$`[\stackrel{~}{L}_n^i,\stackrel{~}{L}_m^j]=\delta _{ij}(nm)\stackrel{~}{L}_{n+m}^i+{\displaystyle \frac{n^3n}{12}}\delta _{ij}\delta _{n,m}.`$
From now on we will simply write $`\stackrel{~}{L}_n^i`$ as $`L_n^i`$.
We introduce operators
$$\mathrm{\Delta }^i=\frac{1}{6}:a^i(z)^3:z^2dz.$$
These are well-defined operators of degree $`0`$ acting on $`S_L`$. One can easily check that
$$\mathrm{\Delta }^i=\frac{1}{2}\underset{n,m>0}{}(a_{nm}^ia_n^ia_m^i+a_n^ia_m^ia_{n+m}^i).$$
Here we omit on the right hand side the terms involving $`a_0`$ since $`a_0`$ acts as $`0`$ on the Fock space $`S_L`$.
###### Lemma 1
We have
$$[\mathrm{\Delta }^i,a_n]=na^i,aL_n^i.$$
proof. It is clear that $`[\mathrm{\Delta }^i,a_n^j]=0`$ for $`ij`$. So it suffices to prove that
$$[\mathrm{\Delta }^i,a_n^i]=nL_n^i.$$
We calculate that
$`[\mathrm{\Delta }^i,a^i(z)]`$ $`=`$ $`\mathrm{Res}_{w=0}{\displaystyle \frac{1}{6}}[:a^i(w)^3:,a^i(z)]w^2`$
$`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{Res}_{w=0}:a^i(w)^2:w^2{\displaystyle \underset{mZ}{}}mw^{m1}z^{m1}`$
$`=`$ $`\mathrm{Res}_{w=0}{\displaystyle \underset{nZ}{}}L_n^iw^{n2}w^2{\displaystyle \underset{mZ}{}}mw^{m1}z^{m1}`$
$`=`$ $`{\displaystyle \underset{nZ}{}}nL_n^iz^{n1},`$
Therefore
$$[\mathrm{\Delta }^i,a_n^i]=\mathrm{Res}_{z=0}[\mathrm{\Delta }^i,a^i(z)]z^n=nL_n^i.$$
$`\mathrm{}`$
## 2 Representation rings of wreath products
Given a finite group $`\mathrm{\Gamma }`$, we denote by $`\mathrm{\Gamma }^{}`$ the set of complex irreducible characters and by $`\mathrm{\Gamma }_{}`$ the set of conjugacy classes. We denote by $`R_Z(\mathrm{\Gamma })`$ the $`Z`$-span of irreducible characters of $`\mathrm{\Gamma }`$. Denote by $`c^0`$ the identity conjugacy class. We identify $`R(\mathrm{\Gamma })=C_ZR_Z(\mathrm{\Gamma })`$ with the space of class functions on $`\mathrm{\Gamma }`$.
For $`c\mathrm{\Gamma }_{}`$ let $`\zeta _c`$ be the order of the centralizer of an element in the class $`c`$. Denote by $`|\mathrm{\Gamma }|`$ the order of $`\mathrm{\Gamma }`$. The usual bilinear form $`,_\mathrm{\Gamma }`$ on $`R(\mathrm{\Gamma })`$ is defined as follows (often we will omit the subscript $`\mathrm{\Gamma }`$):
$`f,g=f,g_\mathrm{\Gamma }={\displaystyle \frac{1}{|\mathrm{\Gamma }|}}{\displaystyle \underset{x\mathrm{\Gamma }}{}}f(x)g(x^1)={\displaystyle \underset{c\mathrm{\Gamma }_{}}{}}\zeta _c^1f(c)g(c^1),`$ (1)
where $`c^1`$ denotes the conjugacy class $`\{x^1,xc\}`$. Clearly $`\zeta _c=\zeta _{c^1}`$. It is well known that
$`\gamma ,\gamma ^{}`$ $`=`$ $`\delta _{\gamma ,\gamma ^{}},\gamma ,\gamma ^{}\mathrm{\Gamma }^{}`$
$`{\displaystyle \underset{\gamma \mathrm{\Gamma }^{}}{}}\gamma (c^{})\gamma (c^1)`$ $`=`$ $`\delta _{c,c^{}}\zeta _c,c,c^{}\mathrm{\Gamma }_{}.`$ (2)
One may regard $`C[\mathrm{\Gamma }]`$ as the space of functions on $`\mathrm{\Gamma }`$, and thus $`R(\mathrm{\Gamma })`$ as a subspace of $`C[\mathrm{\Gamma }]`$. Given $`f,gC[\mathrm{\Gamma }]`$, the convolution $`fgC[\mathrm{\Gamma }]`$ is defined by
$$fg(x)=\underset{y\mathrm{\Gamma }}{}f(xy^1)g(y),f,gC[\mathrm{\Gamma }],x\mathrm{\Gamma }.$$
In particular if $`f,gR(\mathrm{\Gamma })`$, then so is $`fg`$. It is well known that
$`\gamma ^{}\gamma ={\displaystyle \frac{|\mathrm{\Gamma }|}{d_\gamma }}\delta _{\gamma ,\gamma ^{}}\gamma ,\gamma ^{},\gamma \mathrm{\Gamma }^{},`$ (3)
where $`d_\gamma `$ is the degree of the irreducible character $`\gamma `$.
Denote by $`K_c`$ the sum of all elements in the conjugacy class $`c`$. By abuse of notation, we also regard $`K_c`$ the class function on $`\mathrm{\Gamma }`$ which takes value $`1`$ on elements in the conjugacy class $`c`$ and $`0`$ elsewhere. It is clear that $`K_c,c\mathrm{\Gamma }_{},`$ form a basis of $`R(\mathrm{\Gamma })`$. The elements $`K_c,c\mathrm{\Gamma }_{}`$ actually form a linear basis of the center in the group algebra $`C[\mathrm{\Gamma }]`$. But we will not need this fact.
For wreath products we basically follow the excellent presentation of Macdonald \[M\], Appendix B, Chapter 1, with the exception of Theorem 1 which is quoted from \[W\] (also see \[FJW\]). Given a positive integer $`n`$, let $`\mathrm{\Gamma }^n=\mathrm{\Gamma }\times \mathrm{}\times \mathrm{\Gamma }`$ be the $`n`$-th direct product of $`\mathrm{\Gamma }`$. The symmetric group $`S_n`$ acts on $`\mathrm{\Gamma }^n`$ by permutations: $`\sigma (g_1,\mathrm{},g_n)=(g_{\sigma ^1(1)},\mathrm{},g_{\sigma ^1(n)}).`$ The wreath product of $`\mathrm{\Gamma }`$ with $`S_n`$ is defined to be the semi-direct product
$$\mathrm{\Gamma }_n=\{(g,\sigma )|g=(g_1,\mathrm{},g_n)\mathrm{\Gamma }^n,\sigma S_n\}$$
with the multiplication
$$(g,\sigma )(h,\tau )=(g\sigma (h),\sigma \tau ).$$
Let $`\lambda =(\lambda _1,\lambda _2,\mathrm{},\lambda _l)`$ be a partition of integer $`|\lambda |=\lambda _1+\mathrm{}+\lambda _l`$, where $`\lambda _1\mathrm{}\lambda _l1`$. The integer $`l`$ is called the length of the partition $`\lambda `$ and is denoted by $`l(\lambda )`$. We will identify the partition $`(\lambda _1,\lambda _2,\mathrm{},\lambda _l)`$ with $`(\lambda _1,\lambda _2,\mathrm{},\lambda _l,0,\mathrm{},0)`$. We will also make use of another notation for partitions:
$$\lambda =(1^{m_1}2^{m_2}\mathrm{}),$$
where $`m_i`$ is the number of parts in $`\lambda `$ equal to $`i`$.
For a finite set $`X`$ and $`\rho =(\rho (x))_{xX}`$ a family of partitions indexed by $`X`$, we write
$$\rho =\underset{xX}{}|\rho (x)|.$$
Sometimes it is convenient to regard $`\rho =(\rho (x))_{xX}`$ as a partition-valued function on $`X`$. We denote by $`๐ซ(X)`$ the set of all partitions indexed by $`X`$ and by $`๐ซ_n(X)`$ the set of all partitions in $`๐ซ(X)`$ such that $`\rho =n`$.
The conjugacy classes of $`\mathrm{\Gamma }_n`$ can be described in the following way. Let $`x=(g,\sigma )\mathrm{\Gamma }_n`$, where $`g=(g_1,\mathrm{},g_n)\mathrm{\Gamma }^n,`$ $`\sigma S_n`$. The permutation $`\sigma `$ is written as a product of disjoint cycles. For each such cycle $`y=(i_1i_2\mathrm{}i_k)`$ the element $`g_{i_k}g_{i_{k1}}\mathrm{}g_{i_1}\mathrm{\Gamma }`$ is determined up to conjugacy in $`\mathrm{\Gamma }`$ by $`g`$ and $`y`$, and will be called the cycle-product of $`x`$ corresponding to the cycle $`y`$. For any conjugacy class $`c`$ and each integer $`i1`$, the number of $`i`$-cycles in $`\sigma `$ whose cycle-product lies in $`c`$ will be denoted by $`m_i(c)`$. Denote by $`\rho (c)`$ the partition $`(1^{m_1(c)}2^{m_2(c)}\mathrm{})`$, $`c\mathrm{\Gamma }_{}`$. Then each element $`x=(g,\sigma )\mathrm{\Gamma }_n`$ gives rise to a partition-valued function $`(\rho (c))_{c\mathrm{\Gamma }_{}}๐ซ(\mathrm{\Gamma }_{})`$ such that $`_{i,c}im_i(c)=n`$. The partition-valued function $`\rho =(\rho (c))_{cG_{}}`$ is called the type of $`x`$. It is known that any two elements of $`\mathrm{\Gamma }_n`$ are conjugate in $`\mathrm{\Gamma }_n`$ if and only if they have the same type.
Given a partition $`\lambda =(1^{m_1}2^{m_2}\mathrm{})`$, we define $`z_\lambda =_{i1}i^{m_i}m_i!.`$ We note that $`z_\lambda `$ is the order of the centralizer of an element of cycle-type $`\lambda `$ in $`S_{|\lambda |}`$. The order of the centralizer of an element $`x=(g,\sigma )\mathrm{\Gamma }_n`$ of the type $`\rho =(\rho (c))_{c\mathrm{\Gamma }_{}}`$ is
$`Z_\rho ={\displaystyle \underset{c\mathrm{\Gamma }_{}}{}}z_{\rho (c)}\zeta _c^{l(\rho (c))}.`$ (4)
Recall that $`R(\mathrm{\Gamma }_n)=R_Z(\mathrm{\Gamma }_n)_ZC`$. We set
$`R_\mathrm{\Gamma }={\displaystyle \underset{n0}{}}R(\mathrm{\Gamma }_n).`$
A symmetric bilinear form on $`R_\mathrm{\Gamma }`$ is given by
$`u,v={\displaystyle \underset{n0}{}}u_n,v_n_{\mathrm{\Gamma }_n},`$ (5)
where $`u=_nu_n`$ and $`v=_nv_n`$ with $`u_n,v_n\mathrm{\Gamma }_n`$.
Since $`R_Z(\mathrm{\Gamma })`$ may be regarded as an integral lattice with non-degenerate symmetric bilinear form given by (1) and with an orthonormal basis $`\mathrm{\Gamma }^{}`$, we can apply the constructions in Sect. 1 to the lattice $`R_Z(\mathrm{\Gamma })`$. We denote the corresponding Heisenberg algebra by $`\widehat{h}_\mathrm{\Gamma }`$ with generators $`a_n(\gamma ),nZ,\gamma \mathrm{\Gamma }^{}`$, and its irreducible representation by $`S_\mathrm{\Gamma }`$.
By identifying $`a_n(\gamma )`$ $`(n>0,\gamma \mathrm{\Gamma }^{})`$ with the $`n`$-th power sum in a sequence of variables parameterized by $`\gamma \mathrm{\Gamma }^{}`$, we may regard $`S_\mathrm{\Gamma }`$ as the algebra of symmetric functions parameterized by $`\gamma \mathrm{\Gamma }^{}`$. In particular the operator $`a_n(\gamma )`$ $`(n>0,\gamma \mathrm{\Gamma }^{})`$ acts as the differential operator $`n\frac{}{a_n(\gamma )}`$.
For $`mZ,c\mathrm{\Gamma }_{}`$ we define
$`a_m(c)={\displaystyle \underset{\gamma \mathrm{\Gamma }^{}}{}}\gamma (c^1)a_m(\gamma ).`$ (6)
From the orthogonality of the irreducible characters (2) it follows that
$`a_m(\gamma )={\displaystyle \underset{c\mathrm{\Gamma }_{}}{}}\zeta _c^1\gamma (c)a_m(c).`$
Thus $`a_n(c)`$ ($`nZ,c\mathrm{\Gamma }_{}`$) and $`C`$ form a new basis for the Heisenberg algebra $`\widehat{h}_\mathrm{\Gamma }`$.
Given a partition $`\lambda =(\lambda _1,\lambda _2,\mathrm{})`$ and $`c\mathrm{\Gamma }_{}`$, we define
$`p_\lambda (c)=p_{\lambda _1}(c)p_{\lambda _2}(c)\mathrm{}.`$
For any $`\rho =(\rho (c))_{c\mathrm{\Gamma }_{}}๐ซ(\mathrm{\Gamma }_{})`$, we further define
$`P_\rho ={\displaystyle \underset{c\mathrm{\Gamma }_{}}{}}p_{\rho (c)}(c).`$
The elements $`P_\rho ,\rho ๐ซ(\mathrm{\Gamma }_{})`$ form a $`C`$-basis for $`S_\mathrm{\Gamma }`$.
We define a bilinear form $`,`$ on the space $`S_\mathrm{\Gamma }`$ by letting
$`P_\rho ,P_\sigma =\delta _{\rho ,\sigma }Z_\rho .`$ (7)
Let $`\mathrm{\Psi }:\mathrm{\Gamma }_nS_\mathrm{\Gamma }`$ be the map defined by $`\mathrm{\Psi }(x)=P_\rho `$ if $`x\mathrm{\Gamma }_n`$ is of type $`\rho `$. We define a $`C`$-linear map $`\text{ch}:_{n0}C[\mathrm{\Gamma }_n]S_\mathrm{\Gamma }`$ by letting
$`\text{ch}(f)`$ $`=`$ $`{\displaystyle \frac{1}{|\mathrm{\Gamma }_n|}}{\displaystyle \underset{x\mathrm{\Gamma }_n}{}}f(x)\mathrm{\Psi }(x)`$
$`=`$ $`{\displaystyle \underset{\rho ๐ซ(\mathrm{\Gamma }_{})}{}}Z_\rho ^1f_\rho P_\rho \text{if }fR_\mathrm{\Gamma },`$
where $`f_\rho `$ is the value of $`f`$ at elements of type $`\rho `$. Most often we will think of ch as a map from $`R_\mathrm{\Gamma }`$ to $`S_\mathrm{\Gamma }`$ as in \[M\], usually referred to as the characteristic map. It is well known that $`\text{ch}:R_\mathrm{\Gamma }S_\mathrm{\Gamma }`$ is an isometry for the bilinear forms on $`R_\mathrm{\Gamma }`$ and $`S_\mathrm{\Gamma }`$ defined in (5) and (7).
As we identify $`a_n(\gamma )`$ $`(n>0,\gamma \mathrm{\Gamma }^{})`$ with the $`n`$-th power sum $`p_n(\gamma )`$ and the space $`S_\mathrm{\Gamma }`$ with the space of symmetric functions indexed by $`\mathrm{\Gamma }^{}`$, we may regard the Schur function $`s_\lambda (\gamma )`$ associated to $`\gamma `$ and a partition $`\lambda `$ a corresponding element in $`S_\mathrm{\Gamma }`$. For $`\lambda ๐ซ(\mathrm{\Gamma }^{})`$, we denote
$`s_\lambda ={\displaystyle \underset{\gamma \mathrm{\Gamma }^{}}{}}s_{\lambda (\gamma )}(\gamma )S_\mathrm{\Gamma }.`$
Then $`s_\lambda `$ is the image under the characteristic map ch of the character of an irreducible representation $`\chi ^\lambda `$ of $`\mathrm{\Gamma }_n`$ (cf. \[M\]).
Denote by $`c_n(c\mathrm{\Gamma }_{})`$ the conjugacy class in $`\mathrm{\Gamma }_n`$ of elements $`(x,s)\mathrm{\Gamma }_n`$ such that $`s`$ is an $`n`$-cycle and the cycle product of $`x`$ lies in the conjugacy class $`c`$. Denote by $`\sigma _n(c)`$ the class function on $`\mathrm{\Gamma }_n`$ which takes value $`n\zeta _c`$ (i.e. the order of the centralizer of an element in the class $`c_n`$) on elements in the class $`c_n`$ and $`0`$ elsewhere. For $`\rho =\{m_r(c)\}_{r1,c\mathrm{\Gamma }_{}}๐ซ_n(\mathrm{\Gamma }_{})`$, $`\sigma _\rho =_{r1,c\mathrm{\Gamma }_{}}\sigma _r(c)^{m_r(c)}`$ is the class function on $`\mathrm{\Gamma }_n`$ which takes value $`Z_\rho `$ on the conjugacy class of type $`\rho `$ and $`0`$ elsewhere. Given $`\gamma R(\mathrm{\Gamma })`$, we denote by $`\sigma _n(\gamma )`$ the class function on $`\mathrm{\Gamma }_n`$ which takes value $`n\gamma (c)`$ on elements in the class $`c_n,c\mathrm{\Gamma }_{}`$, and $`0`$ elsewhere. We define an operator $`\stackrel{~}{a}_n(\gamma ),n>0`$ to be a map from $`R_\mathrm{\Gamma }`$ to itself by the following composition
$$R(\mathrm{\Gamma }_m)\stackrel{\sigma _n(\gamma )}{}R(\mathrm{\Gamma }_n)R(\mathrm{\Gamma }_m)\stackrel{Ind}{}R(\mathrm{\Gamma }_{n+m}).$$
We also define another operator $`\stackrel{~}{a}_n(\gamma ),n>0`$ to be a map from $`R_\mathrm{\Gamma }`$ to itself (which is the adjoint of $`\stackrel{~}{a}_n(\gamma )`$ with respect to the bilinear form (5) ) as the composition
$$R(\mathrm{\Gamma }_m)\stackrel{Res}{}R(\mathrm{\Gamma }_n)R(\mathrm{\Gamma }_{mn})\stackrel{\sigma _n(\gamma ),}{}R(\mathrm{\Gamma }_{mn}).$$
The following theorem was established in \[W\] (also see \[FJW\]).
###### Theorem 1
The space $`R_\mathrm{\Gamma }`$ affords a representation of the Heisenberg algebra $`\widehat{h}_\mathrm{\Gamma }`$ by letting $`a_n(\gamma )`$ $`(nZ\backslash 0)`$ act as $`\stackrel{~}{a}_n(\gamma )`$, and $`C`$ as $`1`$. The characteristic map ch is an isomorphism of $`R_\mathrm{\Gamma }`$ and $`S_\mathrm{\Gamma }`$ as representations over the Heisenberg algebra.
## 3 Virasoro algebra and group convolution
We first look at the case when $`\mathrm{\Gamma }`$ is trivial and so $`\mathrm{\Gamma }_n`$ becomes the symmetric group $`S_n`$. We simply write the $`i`$-th power sum $`p_i(\gamma )`$ as $`p_i`$.
We consider the convolution product on the space of class functions on $`S_n`$ with the class function $`K_{(1^{n2}2)}`$, which takes value $`1`$ at elements of cycle type $`(1^{n2}2)`$ and $`0`$ otherwise. It follows from (2) that
$$\text{ch}(K_{(1^{n2}2)})=\frac{1}{2(n2)!}p_1^{n2}p_2.$$
We denote
$`\mathrm{\Delta }={\displaystyle \frac{1}{2}}{\displaystyle \underset{i,j1}{}}\left(ijp_{i+j}{\displaystyle \frac{}{p_i}}{\displaystyle \frac{}{p_j}}+(i+j)p_ip_j{\displaystyle \frac{}{p_{i+j}}}\right).`$ (9)
Given $`f,gR(S_n)`$, it is natural to define the convolution of two symmetric functions $`\text{ch}(f)`$ and $`\text{ch}(g)`$ as
$$\text{ch}(f)\text{ch}(g):=\text{ch}(fg).$$
The next theorem describes the effect of the convolution with $`K_{(1^{n2}2)}`$ on the space of symmetric polynomials by means of the characteristic map.
###### Theorem 2
For any $`fC[S_n]`$, we have
$$\text{ch}(K_{(1^{n2}2)}f)=\mathrm{\Delta }\text{ch}(f).$$
Equivalently we have
$$\frac{1}{2(n2)!}p_1^{n2}p_2\text{ch}(f)=\mathrm{\Delta }\text{ch}(f).$$
proof. Take a transposition $`(a,b)`$ and a permutation $`\tau `$.
If $`a`$ and $`b`$ lie in different cycles of $`\tau `$, the product $`(a,b)\tau `$ will have the effect of combining the two cycles, say of cycle length $`i`$ and $`j`$ respectively, in $`\tau `$ containing respectively $`a`$ and $`b`$ into a single cycle. For example, let $`n=7,a=3,b=5`$ and $`\tau =(1,3)(2,6,5)(4,7)`$, then $`(3,5)\tau =(1,5,2,6,3)(4,7)`$. Thus one $`p_ip_j`$ in $`\text{ch}(\tau )`$ is replaced by one $`p_{i+j}`$ in $`\text{ch}((a,b)\tau )`$. Among all transpositions $`(a,b)S_n`$, there are exactly $`ij`$ of which have the effect of replacing one $`p_ip_j`$ by one $`p_{i+j}`$.
On the other hand, if $`a`$ and $`b`$ lie in a same cycle, say of cycle length $`k`$, of $`\tau `$, then the product $`(a,b)\tau `$ will have the effect of splitting this cycle in $`\tau `$ containing $`a`$ and $`b`$ into two disjoint cycles. For example, let $`n=8,a=3,b=5`$ and $`\tau =(8,7,3,1,4,5)(2,6)`$, then $`(3,5)\tau =(8,7,5)(1,4,3)(2,6)`$. More precisely the cycle of $`\tau `$ splits into two disjoint cycles of length $`i`$ and $`ki`$ if $`a`$ and $`b`$ in the cycle of $`\tau `$ are separated by $`i1`$ and $`ki1`$ elements. Thus one $`p_k`$ in $`\text{ch}(\tau )`$ is replaced by one $`p_ip_{ki}`$ in $`\text{ch}((a,b)\tau )`$. We can see easily among all possible transpositions there are $`k`$ (resp. $`k/2`$) of them which have the effect of replacing one $`p_k`$ by one $`p_ip_{ki}`$ when $`k/2i`$ (resp. $`k/2=i`$).
Combining the above considerations together, we have proved the theorem. $`\mathrm{}`$
###### Remark 1
Our purpose of studying this convolution is to find a group theoretic construction of the Virasoro algebra. It turns out that such a convolution appeared earlier in a paper of Goulden (\[Go\], Proposition 3.1) in his study of the number of ways of writing permutations in a given conjugacy class as products of transpositions.
Recall that $`s_\lambda `$ denotes the Schur functions associated to the partition $`\lambda `$. We denote by $`f^\lambda `$ the degree of $`s_\lambda `$, i.e. the dimension of the irreducible representation of $`S_n`$ corresponding to $`s_\lambda `$.
###### Proposition 2
We have
$`\mathrm{\Delta }(s_\lambda )={\displaystyle \frac{n(n1)}{2f^\lambda }}p_1^{n2}p_2,s_\lambda s_\lambda .`$
Proof. By the orthogonality relation of characters (2) we have $`s_\mu ,s_\lambda =\delta _{\mu ,\lambda }`$, and thus $`p_1^{n2}p_2=_\mu p_1^{n2}p_2,s_\mu s_\mu .`$ In addition Eq. (3) implies that $`s_\mu s_\lambda =\frac{n!}{f^\lambda }\delta _{\mu ,\lambda }s_\lambda .`$ Thus by Theorem 2, we have
$`\mathrm{\Delta }(s_\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{2(n2)!}}p_1^{n2}p_2s_\lambda `$
$`=`$ $`{\displaystyle \frac{1}{2(n2)!}}{\displaystyle \underset{\mu }{}}p_1^{n2}p_2,s_\mu s_\mu s_\lambda `$
$`=`$ $`{\displaystyle \frac{n(n1)}{2f^\lambda }}p_1^{n2}p_2,s_\lambda s_\lambda .`$
$`\mathrm{}`$
Now we return to the general case of wreath product $`\mathrm{\Gamma }_n`$. Given $`c\mathrm{\Gamma }_{}`$, we denote by $`\lambda _c๐ซ_n(\mathrm{\Gamma }_{})`$ the function which maps $`c`$ to the one-part partition $`(2)`$, $`c^0`$ to the partition $`(1^{n2})`$ and other conjugacy classes to $`0`$. We denote by $`K_{\lambda _c}`$ the sum of all the elements in the conjugacy class corresponding to $`\lambda _c`$.
We introduce the operator acting on $`R_\mathrm{\Gamma }`$
$$\mathrm{\Delta }^\gamma =\frac{1}{6}:\stackrel{~}{a}^\gamma (z)^3:z^2dz$$
where $`\stackrel{~}{a}^\gamma (z)=_{nZ}\stackrel{~}{a}_n(\gamma )z^{n1}.`$
We define
$`\mathrm{\Delta }_c={\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}{\displaystyle \frac{|\mathrm{\Gamma }|^2\beta (c^1)}{d_\beta ^2\zeta _c}}\mathrm{\Delta }^\beta .`$ (10)
The operator $`\mathrm{\Delta }_c`$ reduces to (9) for $`\mathrm{\Gamma }`$ trivial. By the characteristic map, the operator $`\mathrm{\Delta }^\gamma `$ can be identified with the differential operator
$$\frac{1}{2}\underset{i,j1}{}\left(ijp_{i+j}(\gamma )\frac{}{p_i(\gamma )}\frac{}{p_j(\gamma )}+(i+j)p_i(\gamma )p_j(\gamma )\frac{}{p_{i+j}(\gamma )}\right).$$
Note that
$$\text{ch}(K_{\lambda _c})=\frac{1}{Z_{\lambda _c}}p_1(c^0)^{n2}p_2(c),$$
where
$`Z_{\lambda _c}=2(n2)!|\mathrm{\Gamma }|^{n2}\zeta _c`$ (11)
is the order of the centralizer of an element in the conjugacy class associated to $`\lambda _c`$, cf. (4). The following theorem generalizes Theorem 2.
###### Theorem 3
Given $`fR_\mathrm{\Gamma }`$, we have
$$\text{ch}(K_{\lambda _c}f)=\mathrm{\Delta }_c\text{ch}(f).$$
Equivalently we have
$$\frac{1}{Z_{\lambda _c}}p_2(c)p_1(c^0)^{n2}\text{ch}(f)=\mathrm{\Delta }_c\text{ch}(f).$$
We need some preparation for the proof of the theorem. Recall that the image of the irreducible character associated to $`\lambda ๐ซ_n(\mathrm{\Gamma }^{})`$ under the characteristic map is $`s_\lambda =_\gamma s_{\lambda (\gamma )}(\gamma )`$. We set $`n_\gamma =|\lambda (\gamma )|`$. The first lemma below is straightforward.
###### Lemma 2
Given $`\beta \mathrm{\Gamma }^{}`$ and a sequence of non-negative integers $`m_\gamma ,\gamma \mathrm{\Gamma }^{}`$ such that $`m_\gamma n_\gamma `$ for at least one $`\gamma `$, we have
$`p_1(\beta )^{m_\beta 2}p_2(\beta ){\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{m_\gamma },s_\lambda =0.`$
###### Lemma 3
Given $`\beta \mathrm{\Gamma }^{}`$, $`\lambda ๐ซ_n(\mathrm{\Gamma }^{})`$, and setting $`n_\gamma =|\lambda (\gamma )|`$ (for all $`\gamma \mathrm{\Gamma }^{}`$), we have
$`{\displaystyle \frac{n_\beta (n_\beta 1)}{2}}p_1(\beta )^{n_\beta 2}p_2(\beta ){\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma },s_\lambda s_\lambda `$
$`=`$ $`\left({\displaystyle \underset{\gamma }{}}f^{\lambda (\gamma )}\right)\mathrm{\Delta }^\beta s_{\lambda (\beta )}(\beta ){\displaystyle \underset{\gamma \beta }{}}s_{\lambda (\gamma )}(\gamma ).`$
Proof. It follows from Proposition 2 that
$`\mathrm{\Delta }^\beta (s_{\lambda (\beta )}(\beta ))={\displaystyle \frac{n_\beta (n_\beta 1)}{2f^{\lambda (\beta )}}}p_1(\beta )^{n_\beta 2}p_2(\beta ),s_{\lambda (\beta )}(\beta )s_{\lambda (\beta )}(\beta ).`$ (12)
Noting also that
$`p_1(\gamma )^{n_\gamma },s_{\lambda (\gamma )}(\gamma )=f^{\lambda (\gamma )},`$ (13)
we calculate that
$`{\displaystyle \frac{n_\beta (n_\beta 1)}{2}}p_1(\beta )^{n_\beta 2}p_2(\beta ){\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma },s_\lambda s_\lambda `$
$`=`$ $`{\displaystyle \frac{n_\beta (n_\beta 1)}{2}}p_1(\beta )^{n_\beta 2}p_2(\beta ),s_{\lambda (\beta )}(\beta )s_{\lambda (\beta )}(\beta )`$
$`{\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma },s_{\lambda (\gamma )}(\gamma )s_{\lambda (\gamma )}(\gamma )`$
$`=`$ $`f^{\lambda (\beta )}\mathrm{\Delta }^\beta s_{\lambda (\beta )}(\beta ){\displaystyle \underset{\gamma \beta }{}}f^{\lambda (\gamma )}s_{\lambda (\gamma )}(\gamma )\text{ by Eqs. }(\text{12})\text{ and }(\text{13}),`$
$`=`$ $`\left({\displaystyle \underset{\gamma \mathrm{\Gamma }^{}}{}}f^{\lambda (\gamma )}\right)\mathrm{\Delta }^\beta s_{\lambda (\beta )}(\beta ){\displaystyle \underset{\gamma \beta }{}}s_{\lambda (\gamma )}(\gamma ).`$
This proves the lemma. $`\mathrm{}`$
Proof of Theorem 3 . Consider the irreducible character associated to $`\lambda ๐ซ_n(\mathrm{\Gamma }^{})`$ which maps by the characteristic map ch to $`s_\lambda =_\gamma s_{\lambda (\gamma )}(\gamma )`$. We set $`n_\gamma =|\lambda (\gamma )|`$.
We calculate that
$`{\displaystyle \frac{1}{2(n2)!}}p_1(c^0)^{n2}p_2(c)`$ (14)
$`=`$ $`{\displaystyle \frac{1}{2(n2)!}}\left({\displaystyle \underset{\gamma \mathrm{\Gamma }^{}}{}}d_\gamma p_1(\gamma )\right)^{n2}\left({\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}\beta (c^1)p_2(\beta )\right)\text{ by Eq.}(\text{6}),`$
$`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \underset{\{m_\gamma \}}{}}{\displaystyle \underset{\gamma \mathrm{\Gamma }^{}}{}}{\displaystyle \frac{(d_\gamma p_1(\gamma ))^{m_\gamma }}{m_\gamma !}}\right)\left({\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}\beta (c^1)p_2(\beta )\right)`$
$`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{\{m_\gamma \}}{}}{\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}\beta (c^1)p_2(\beta ){\displaystyle \frac{(d_\beta p_1(\beta ))^{m_\beta }}{(m_\beta )!}}{\displaystyle \underset{\gamma \beta }{}}{\displaystyle \frac{(d_\gamma p_1(\gamma ))^{m_\gamma }}{m_\gamma !}},`$
where the sum $`_{\{m_\gamma \}}`$ is taken over all sequences of nonnegative integers $`m_\gamma ,\gamma \mathrm{\Gamma }^{}`$ such that $`_\gamma m_\gamma =n2`$.
Recall \[M, Z\] that the degree of the irreducible character associated to $`\lambda ๐ซ_n(\mathrm{\Gamma }^{})`$ is
$`f^\lambda =n!{\displaystyle \underset{\gamma }{}}d_\gamma ^{n_\gamma }f^{\lambda (\gamma )}/n_\gamma !.`$ (15)
We also obtain by using (3) and $`|\mathrm{\Gamma }_n|=n!|\mathrm{\Gamma }|^n`$ that
$`{\displaystyle \frac{f^\lambda }{n!|\mathrm{\Gamma }|^n}}\left(p_2(\beta )p_1(\beta )^{n_\beta 2}{\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma }\right)s_\lambda `$ (16)
$`=`$ $`p_2(\beta )p_1(\beta )^{n_\beta 2}{\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma },s_\lambda s_\lambda .`$
We then have
$`{\displaystyle \frac{1}{Z_{\lambda _c}}}p_1(c^0)^{n2}p_2(c)s_\lambda `$
$`=`$ $`{\displaystyle \frac{1}{2|\mathrm{\Gamma }|^{n2}\zeta _c}}{\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}\beta (c^1)p_2(\beta ){\displaystyle \frac{(d_\beta p_1(\beta ))^{n_\beta 2}}{(n_\beta 2)!}}{\displaystyle \underset{\gamma \beta }{}}{\displaystyle \frac{(d_\gamma p_1(\gamma ))^{n_\gamma }}{n_\gamma !}}s_\lambda `$
$`\text{ by Eqs.}(\text{11}),(\text{14}),\text{ and Lemma}\text{2},`$
$`=`$ $`{\displaystyle \frac{1}{|\mathrm{\Gamma }|^{n2}\zeta _c}}{\displaystyle \underset{\gamma }{}}{\displaystyle \frac{d_\gamma ^{n_\gamma }}{n_\gamma !}}{\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}{\displaystyle \frac{n_\beta (n_\beta 1)\beta (c^1)}{2d_\beta ^2}}\left(p_2(\beta )p_1(\beta )^{n_\beta 2}{\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma }\right)s_\lambda `$
$`=`$ $`{\displaystyle \frac{f^\lambda }{n!|\mathrm{\Gamma }|^{n2}\zeta _c}}{\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}{\displaystyle \frac{n_\beta (n_\beta 1)\beta (c^1)}{2d_\beta ^2(_\gamma f^{\lambda (\gamma )})}}\left(p_2(\beta )p_1(\beta )^{n_\beta 2}{\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma }\right)s_\lambda `$
$`\text{ by Eq.}(\text{15}),`$
$`=`$ $`{\displaystyle \frac{|\mathrm{\Gamma }|^2}{\zeta _c}}{\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}{\displaystyle \frac{n_\beta (n_\beta 1)\beta (c^1)}{2d_\beta ^2(_\gamma f^{\lambda (\gamma )})}}p_2(\beta )p_1(\beta )^{n_\beta 2}{\displaystyle \underset{\gamma \beta }{}}p_1(\gamma )^{n_\gamma },s_\lambda s_\lambda `$
$`\text{ by Eq.}(\text{16}),`$
$`=`$ $`{\displaystyle \underset{\beta \mathrm{\Gamma }^{}}{}}{\displaystyle \frac{|\mathrm{\Gamma }|^2\beta (c^1)}{d_\beta ^2\zeta _c}}\left(\mathrm{\Delta }^\beta s_{\lambda (\beta )}(\beta ){\displaystyle \underset{\gamma \beta }{}}s_{\lambda (\gamma )}(\gamma )\right)`$
$`\text{ by Lemma}\text{3},`$
$`=`$ $`\mathrm{\Delta }_cs_\lambda .`$
The last equation holds since $`\mathrm{\Delta }^\beta s_{\lambda (\gamma )}(\gamma )=0`$ for $`\gamma \beta `$. This finishes the proof, since $`s_\lambda ,\lambda ๐ซ(\mathrm{\Gamma }^{})`$ form a basis of the space $`R_\mathrm{\Gamma }`$. $`\mathrm{}`$
###### Corollary 1
For the identity conjugacy class $`c^0\mathrm{\Gamma }_{}`$, we have $`\mathrm{\Delta }_{c^0}=_{\beta \mathrm{\Gamma }^{}}\frac{|\mathrm{\Gamma }|}{d_\beta }\mathrm{\Delta }^\beta .`$
Combining (10), Lemma 1, Theorem 1 and Theorem 3, we obtain the following.
###### Theorem 4
The operator $`\mathrm{\Delta }_c`$ acting on $`R_\mathrm{\Gamma }`$ is realized by the convolution with $`K_{\lambda _c}`$. The commutation relation between $`\mathrm{\Delta }_c`$ and the Heisenberg algebra generator $`\stackrel{~}{a}_n(\gamma )`$ (constructed in a group theoretic manner) is given by
$$[\mathrm{\Delta }_c,\stackrel{~}{a}_n(\gamma )]=\frac{n|\mathrm{\Gamma }|^2\gamma (c^1)}{d_\gamma ^2\zeta _c}L_n^\gamma ,$$
where the operators $`L_n^\gamma `$ acting on the space $`R_\mathrm{\Gamma }`$ satisfy the Virasoro commutation relations:
$$[L_n^\beta ,L_m^\gamma ]=\delta _{\beta \gamma }L_{n+m}^\gamma +\frac{n^3n}{12}\delta _{\beta \gamma }\delta _{n,m}.$$
Frenkel: Department of Mathematics, Yale University, New Haven, CT 06520; Wang: Department of Mathematics, North Carolina State University, Raleigh, NC 27695, wqwang@math.ncsu.edu.
|
warning/0006/cond-mat0006077.html
|
ar5iv
|
text
|
# Influence of nonlocal electrodynamics on the anisotropic vortex pinning in ๐โข๐โข๐โโข๐ตโโข๐ถ
## I References
Fig.1: Pinning force density $`F_p`$ vs $`H`$ for $`YNi_2B_2C`$ single crystal with (a) $`๐๐`$-axis and (b) $`๐`$ (open symbols) and $`๐`$ (solid symbols) at several temperatures.
Fig.2: Reduced pinning force density $`F_p/F_p^{max}`$ vs $`H/H_{c2}`$ at several temperatures with $`๐๐`$-axis. The inset shows $`F_p`$ versus $`H`$ for $`๐c`$ ($`\mathrm{\Theta }=0^{}`$) and tilted $`10^{}`$ and $`30^{}`$ from $`c`$.
Fig.3: Upper panel: the normal state magnetization $`M_{ns}`$ at $`T=7K`$ and $`H=45kOe`$, as a function of the angle $`\phi `$ between the applied field $`๐`$ (contained in the $`\mathrm{๐๐}`$ basal plane) and the $`a`$ axis. Lower panel: pinning force density $`F_p(\phi )`$ at $`T=7K`$ and several fields. Inset: oscillation amplitude of the in-plane normal state magnetization, $`\delta M_{ns}=M_{ns}[110]M_{ns}[100]`$, vs $`H/T`$.
|
warning/0006/hep-th0006040.html
|
ar5iv
|
text
|
# Untitled Document
QUANTUM FIELD THEORY FROM
FIRST PRINCIPLES
Giampiero Esposito Istituto Nazionale di Fisica Nucleare, Sezione di Napoli, Complesso Universitario di Monte S. Angelo, Via Cintia, Edificio Nโ, 80126 Napoli, Italy Dipartimento di Scienze Fisiche, Complesso Universitario di Monte S. Angelo, Via Cintia, Edificio Nโ, 80126 Napoli, Italy
Abstract. When quantum fields are studied on manifolds with boundary, the corresponding one-loop quantum theory for bosonic gauge fields with linear covariant gauges needs the assignment of suitable boundary conditions for elliptic differential operators of Laplace type. There are however deep reasons to modify such a scheme and allow for pseudo-differential boundary-value problems. When the boundary operator is allowed to be pseudo-differential while remaining a projector, the conditions on its kernel leading to strong ellipticity of the boundary-value problem are studied in detail. This makes it possible to develop a theory of one-loop quantum gravity from first principles only, i.e. the physical principle of invariance under infinitesimal diffeomorphisms and the mathematical requirement of a strongly elliptic theory. It therefore seems that a non-local formulation of quantum field theory has some attractive features which deserve further investigation.
The space-time approach to quantum mechanics and quantum field theory has led to several profound developments in the understanding of quantum theory and space-time structure at very high energies.<sup>1,2</sup> In particular, we are here concerned with the choice of boundary conditions. On using path integrals, which lead, in principle, to the appropriate formulation of the ideas of Feynman, DeWitt and many other authors,<sup>2-5</sup> the assignment of boundary conditions consists of two main steps: (i) Choice of Riemannian geometries and field configurations to be included in the path-integral representation of transition amplitudes. (ii) Choice of boundary data to be imposed on the hypersurfaces $`\mathrm{\Sigma }_1`$ and $`\mathrm{\Sigma }_2`$ bounding the given space-time region. The main object of our investigation is the second problem of such a list, when a one-loop approximation is studied for a bosonic gauge theory in linear covariant gauges. The well posed mathematical formulation relies on the โEuclidean approachโ, i.e., in geometric language, on the use of differentiable manifolds endowed with positive-definite metrics $`g`$, so that Lorentzian space-time is actually replaced by an $`m`$-dimensional Riemannian manifold $`(M,g)`$.
In particular, in Euclidean quantum gravity, mixed boundary conditions on metric perturbations $`h_{cd}`$ occur naturally if one requires their complete invariance under infinitesimal diffeomorphisms, as is proved in detail in Ref. 6. On denoting by $`N^a`$ the inward-pointing unit normal to the boundary, by
$$q_b^a\delta _b^aN^aN_b$$
$`(1)`$
the projector of tensor fields onto $`M`$, with associated projection operator
$$\mathrm{\Pi }_{ab}^{cd}q_{(a}^cq_{b)}^d,$$
$`(2)`$
the gauge-invariant boundary conditions for one-loop quantum gravity read<sup>6</sup>
$$\left[\mathrm{\Pi }_{ab}^{cd}h_{cd}\right]_M=0,$$
$`(3)`$
$$\left[\mathrm{\Phi }_a(h)\right]_M=0,$$
$`(4)`$
where $`\mathrm{\Phi }_a`$ is the gauge-averaging functional necessary to obtain an invertible operator $`P_{ab}^{cd}`$ on metric perturbations. When $`P_{ab}^{cd}`$ is chosen to be of Laplace type, $`\mathrm{\Phi }_a`$ reduces to the familiar de Donder term
$$\mathrm{\Phi }_a(h)=^b(h_{ab}\frac{1}{2}g_{ab}g^{cd}h_{cd})=E_a^{bcd}_bh_{cd},$$
$`(5)`$
where $`E^{abcd}`$ is the DeWitt supermetric on the vector bundle of symmetric rank-two tensor fields over $`M`$ ($`g`$ being the metric on $`M`$):
$$E^{abcd}\frac{1}{2}(g^{ac}g^{bd}+g^{ad}g^{bc}g^{ab}g^{cd}).$$
$`(6)`$
The boundary conditions (3) and (4) can then be cast in the GrubbโGilkeyโSmith form:<sup>7,8</sup>
$$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \mathrm{\Lambda }& I\mathrm{\Pi }\end{array}\right)\left(\begin{array}{c}[\phi ]_M\\ [\phi _{;N}]_M\end{array}\right)=0.$$
$`(7)`$
However, the work in Ref. 6 has shown that an operator of Laplace type on metric perturbations is then incompatible with the requirement of strong ellipticity of the boundary-value problem, because the operator $`\mathrm{\Lambda }`$ contains tangential derivatives of metric perturbations.
To take care of this serious drawback, the work in Ref. 9 has proposed to consider in the boundary condition (4) a gauge-averaging functional given by the de Donder term (5) plus an integro-differential operator on metric perturbations, i.e.
$$\mathrm{\Phi }_a(h)E_a^{bcd}_bh_{cd}+_M\zeta _a^{cd}(x,x^{})h_{cd}(x^{})๐V^{}.$$
$`(8)`$
We now point out that the resulting boundary conditions can be cast in the form
$$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }}& I\mathrm{\Pi }\end{array}\right)\left(\begin{array}{c}[\phi ]_M\\ [\phi _{;N}]_M\end{array}\right)=0,$$
$`(9)`$
where $`\stackrel{~}{\mathrm{\Lambda }}`$ reflects the occurrence of the integral over $`M`$ in Eq. (8). It is convenient to work first in a general way and then consider the form taken by these operators in the gravitational case. On requiring that the resulting boundary operator
$$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }}& I\mathrm{\Pi }\end{array}\right)$$
$`(10)`$
should remain a projector: $`^2=`$, we find the condition
$$(\mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }})\mathrm{\Pi }\mathrm{\Pi }(\mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }})=0,$$
$`(11)`$
which reduces to
$$\mathrm{\Pi }\stackrel{~}{\mathrm{\Lambda }}=\stackrel{~}{\mathrm{\Lambda }}\mathrm{\Pi },$$
$`(12)`$
by virtue of the property $`\mathrm{\Pi }\mathrm{\Lambda }=\mathrm{\Lambda }\mathrm{\Pi }=0`$ considered in Ref. 6.
In Euclidean quantum gravity at one-loop level, Eq. (12) leads to
$$\mathrm{\Pi }_{ac}^{br}(x)_M\zeta _b^{cq}(x,x^{})h_{qr}(x^{})๐V^{}=_M\zeta _a^{cd}(x,x^{})\mathrm{\Pi }_{cd}^{qr}(x^{})h_{qr}(x^{})๐V^{},$$
$`(13)`$
which can be re-expressed in the form
$$_M\left[\mathrm{\Pi }_{ac}^{br}(x)\zeta _b^{cq}(x,x^{})\zeta _a^{cd}(x,x^{})\mathrm{\Pi }_{cd}^{qr}(x^{})\right]h_{qr}(x^{})๐V^{}=0.$$
$`(14)`$
Since this should hold for all $`h_{qr}(x^{})`$, it eventually leads to the vanishing of the term in square brackets in the integrand. The notation $`\zeta _b^{cq}(x,x^{})`$ is indeed rather awkward, because there is an even number of arguments, i.e. $`x`$ and $`x^{}`$, with an odd number of indices. Hereafter, we therefore assume that a vector field $`T`$ and kernel $`\stackrel{~}{\zeta }`$ exist such that
$$\zeta _b^{cq}(x,x^{})T^p(x)\stackrel{~}{\zeta }_{bp}^{cq}(x,x^{})T^p\stackrel{~}{\zeta }_{bp}^{c^{}q^{}}.$$
$`(15)`$
The projector condition (12) is therefore satisfied if and only if<sup>10</sup>
$$T^p(x)\left[\mathrm{\Pi }_{ac}^{br}(x)\stackrel{~}{\zeta }_{bp}^{cq}(x,x^{})\stackrel{~}{\zeta }_{ap}^{cd}(x,x^{})\mathrm{\Pi }_{cd}^{qr}(x^{})\right]=0.$$
$`(16)`$
We are now concerned with the issue of ellipticity of the boundary-value problem of one-loop quantum gravity. For this purpose, we begin by recalling what is known about ellipticity of the Laplacian (hereafter $`P`$) on a Riemannian manifold with smooth boundary. This concept is studied in terms of the leading symbol of $`P`$. It is indeed well known that the Fourier transform makes it possible to associate to a differential operator of order $`k`$ a polynomial of degree $`k`$, called the characteristic polynomial or symbol. The leading symbol, $`\sigma _L`$, picks out the highest order part of this polynomial. For the Laplacian, it reads
$$\sigma _L(P;x,\xi )=|\xi |^2I=g^{\mu \nu }\xi _\mu \xi _\nu I.$$
$`(17)`$
With a standard notation, $`(x,\xi )`$ are local coordinates for $`T^{}(M)`$, the cotangent bundle of $`M`$. The leading symbol of $`P`$ is trivially elliptic in the interior of $`M`$, since the right-hand side of (17) is positive-definite, and one has
$$\mathrm{det}[\sigma _L(P;x,\xi )\lambda ]=(|\xi |^2\lambda )^{\mathrm{dim}V}0,$$
$`(18)`$
for all $`\lambda ๐๐_+`$. In the presence of a boundary, however, one needs a more careful definition of ellipticity. First, for a manifold $`M`$ of dimension $`m`$, the $`m`$ coordinates $`x`$ are split into $`m1`$ local coordinates on $`M`$, hereafter denoted by $`\left\{\widehat{x}^k\right\}`$, and $`r`$, the geodesic distance to the boundary. Moreover, the $`m`$ coordinates $`\xi _\mu `$ are split into $`m1`$ coordinates $`\left\{\zeta _j\right\}`$ (with $`\zeta `$ being a cotangent vector on the boundary), jointly with a real parameter $`\omega T^{}(๐)`$. At a deeper level, all this reflects the split
$$T^{}(M)=T^{}(M)T^{}(๐)$$
$`(19)`$
in a neighbourhood of the boundary.<sup>6,11</sup>
The ellipticity we are interested in requires now that $`\sigma _L`$ should be elliptic in the interior of $`M`$, as specified before, and that strong ellipticity should hold. This means that a unique solution exists of the differential equation obtained from the leading symbol:
$$[\sigma _L(P;\left\{\widehat{x}^k\right\},r=0,\left\{\zeta _j\right\},\omega i\frac{}{r})\lambda ]\phi (r,\widehat{x},\zeta ;\lambda )=0,$$
$`(20)`$
subject to the boundary conditions
$$\sigma _g(B)(\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\})\psi (\phi )=\psi ^{}(\phi )$$
$`(21)`$
and to the asymptotic condition
$$\underset{r\mathrm{}}{lim}\phi (r,\widehat{x},\zeta ;\lambda )=0.$$
$`(22)`$
In Eq. (21), $`\sigma _g`$ is the graded leading symbol of the boundary operator in the local coordinates $`\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\}`$, and is given by
$$\sigma _g(B)=\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ i\mathrm{\Gamma }^j\zeta _j& I\mathrm{\Pi }\end{array}\right).$$
$`(23)`$
Roughly speaking, the above construction uses Fourier transform and the inward geodesic flow to obtain the ordinary differential equation (20) from the Laplacian, with corresponding Fourier transform (21) of the original boundary conditions. The asymptotic condition (22) picks out the solutions of Eq. (20) which satisfy Eq. (21) with arbitrary boundary data $`\psi ^{}(\phi )C^{\mathrm{}}(W^{},M)`$ for $`W^{}`$ a vector bundle over the boundary, and vanish at infinite geodesic distance to the boundary. When all the above conditions are satisfied $`\zeta T^{}(M),\lambda ๐๐_+,(\zeta ,\lambda )(0,0)`$ and $`\psi ^{}(\phi )C^{\mathrm{}}(W^{},M)`$, the boundary-value problem $`(P,B)`$ for the Laplacian is said to be strongly elliptic with respect to the cone $`๐๐_+`$.
However, when the gauge-averaging functional (8) is used in the boundary condition (4), the work in Ref. 9 has proved that the operator on metric perturbations takes the form of an operator of Laplace type $`P_{ab}^{cd}`$ plus an integral operator $`G_{ab}^{cd}`$. Explicitly, one finds<sup>9</sup> (with $`R_{bcd}^a`$ being the Riemann curvature of the background geometry $`(M,g)`$)
$$P_{ab}^{cd}=E_{ab}^{cd}(\text{ }\text{ / }\text{ }+R)2E_{ab}^{qf}R_{qpf}^cg^{dp}E_{ab}^{pd}R_p^cE_{ab}^{cp}R_p^d,$$
$`(24)`$
$$G_{ab}^{cd}=U_{ab}^{cd}+V_{ab}^{cd},$$
$`(25)`$
where
$$U_{ab}^{cd}h_{cd}(x)=2E_{rsab}^r_MT^p(x)\stackrel{~}{\zeta }_p^{scd}(x,x^{})h_{cd}(x^{})๐V^{},$$
$`(26)`$
$$h^{ab}V_{ab}^{cd}h_{cd}(x)=_{M^2}h^{ab}(x^{})T^q(x)\stackrel{~}{\zeta }_{pqab}(x,x^{})T^r(x)\stackrel{~}{\zeta }_r^{pcd}(x,x^{\prime \prime })h_{cd}(x^{\prime \prime })๐V^{}๐V^{\prime \prime }.$$
$`(27)`$
We now assume that the operator on metric perturbations, which is so far an integro-differential operator defined by a kernel, is also pseudo-differential. This means that it can be characterized by suitable regularity properties obeyed by the symbol. More precisely, let $`S^d`$ be the set of all symbols $`p(x,\xi )`$ such that (1) $`p`$ is $`C^{\mathrm{}}`$ in $`(x,\xi )`$, with compact $`x`$ support. (2) For all $`(\alpha ,\beta )`$, there exist constants $`C_{\alpha ,\beta }`$ for which
$$\begin{array}{ccc}& \left|(i)^{_{k=1}^m(\alpha _k+\beta _k)}\left(\frac{}{x_1}\right)^{\alpha _1}\mathrm{}\left(\frac{}{x_m}\right)^{\alpha _m}\left(\frac{}{\xi _1}\right)^{\beta _1}\mathrm{}\left(\frac{}{\xi _m}\right)^{\beta _m}p(x,\xi )\right|\hfill & \\ & C_{\alpha ,\beta }\left(1+\sqrt{g^{ab}(x)\xi _a\xi _b}\right)^{d_{k=1}^m\beta _k},\hfill & (28)\hfill \end{array}$$
for some real (not necessarily positive) value of $`d`$. The associated pseudo-differential operator, defined on the Schwarz space and taking values in the set of smooth functions on $`M`$ with compact support:
$$P:๐ฎC_c^{\mathrm{}}(M)$$
acts according to
$$Pf(x)e^{i(xy)\xi }p(x,\xi )f(y)\mu (y,\xi ),$$
$`(29)`$
where $`\mu (y,\xi )`$ is here meant to be the invariant integration measure with respect to $`y_1,\mathrm{},y_m`$ and $`\xi _1,\mathrm{},\xi _m`$. Actually, one first gives the definition for pseudo-differential operators $`P:๐ฎC_c^{\mathrm{}}(๐^m)`$, eventually proving that a coordinate-free definition can be given and extended to smooth Riemannian manifolds.<sup>11</sup>
In the presence of pseudo-differential operators, both ellipticity in the interior of $`M`$ and strong ellipticity of the boundary-value problem need a more involved formulation. In our paper, inspired by the flat-space analysis in Ref. 12, we make the following requirements.<sup>10</sup>
(i) Ellipticity in the Interior
Let $`U`$ be an open subset with compact closure in $`M`$, and consider an open subset $`U_1`$ whose closure $`\overline{U}_1`$ is properly included into $`U`$: $`\overline{U}_1U`$. If $`p`$ is a symbol of order $`d`$ on $`U`$, it is said to be elliptic on $`U_1`$ if there exists an open set $`U_2`$ which contains $`\overline{U}_1`$ and positive constants $`C_0,C_1`$ so that
$$|p(x,\xi )|^1C_1(1+|\xi |)^d,$$
$`(30)`$
for $`|\xi |C_0`$ and $`xU_2`$, where $`|\xi |\sqrt{g^{ab}(x)\xi _a\xi _b}`$. The corresponding operator $`P`$ is then elliptic.
(ii) Strong Ellipticity in the Absence of Boundaries
Let us assume that the symbol under consideration is polyhomogeneous, in that it admits an asymptotic expansion of the form
$$p(x,\xi )\underset{l=0}{\overset{\mathrm{}}{}}p_{dl}(x,\xi ),$$
$`(31)`$
where each term $`p_{dl}`$ has the homogeneity property
$$p_{dl}(x,t\xi )=t^{dl}p_{dl}(x,\xi )\mathrm{if}t1\mathrm{and}|\xi |1.$$
$`(32)`$
The leading symbol is then, by definition,
$$p^0(x,\xi )p_d(x,\xi ).$$
$`(33)`$
Strong ellipticity in the absence of boundaries is formulated in terms of the leading symbol, and it requires that
$$\mathrm{Re}p^0(x,\xi )c(x)|\xi |^d,$$
$`(34)`$
where $`xM`$ and $`|\xi |1`$, $`c`$ being a positive function on $`M`$. It can then be proved that the Gรคrding inequality holds, according to which, for any $`\epsilon >0`$,
$$\mathrm{Re}(Pu,u)bu_{\frac{d}{2}}^2b_1u_{\frac{d}{2}\epsilon }^2\mathrm{for}uH^{\frac{d}{2}}(M),$$
$`(35)`$
with $`b>0`$.
(iii) Strong Ellipticity in the Presence of Boundaries
The homogeneity property (32) only holds for $`t1`$ and $`|\xi |1`$. Consider now the case $`l=0`$, for which one obtains the leading symbol which plays the key role in the definition of ellipticity. If $`p^0(x,\xi )p_d(x,\xi )\sigma _L(P;x,\xi )`$ is not a polynomial (which corresponds to the genuinely pseudo-differential case) while being a homogeneous function of $`\xi `$, it is irregular at $`\xi =0`$. When $`|\xi |1`$, the only control over the leading symbol is provided by estimates of the form<sup>12</sup>
$$\begin{array}{ccc}& |(i)^{_{k=1}^m(\alpha _k+\beta _k)}\left(\frac{}{x_1}\right)^{\alpha _1}...\left(\frac{}{x_m}\right)^{\alpha _m}\left(\frac{}{\xi _1}\right)^{\beta _1}\mathrm{}\left(\frac{}{\xi _m}\right)^{\beta _m}p^0(x,\xi )|\hfill & \\ & c(x)\xi ^{d|\beta |}.\hfill & (36)\hfill \end{array}$$
We therefore come to appreciate the problematic aspect of symbols of pseudo-differential operators.<sup>12</sup> The singularity at $`\xi =0`$ can be dealt with either by modifying the leading symbol for small $`\xi `$ to be a $`C^{\mathrm{}}`$ function (at the price of loosing the homogeneity there), or by keeping the strict homogeneity and dealing with the singularity at $`\xi =0.^{12}`$
On the other hand, we are interested in a definition of strong ellipticity of pseudo-differential boundary-value problems that reduces to Eqs. (20)โ(22) when both $`P`$ and the boundary operator reduce to the form considered in Ref. 6. For this purpose, and bearing in mind the occurrence of singularities in the leading symbols of $`P`$ and of the boundary operator, we make the following requirements.<sup>10</sup>
Let $`(P+G)`$ be a pseudo-differential operator subject to boundary conditions described by the pseudo-differential boundary operator $``$ (the consideration of $`(P+G)`$ rather than only $`P`$ is necessary to achieve self-adjointness, as is described in detail in Refs. 12 and 13). The pseudo-differential boundary-value problem $`((P+G),)`$ is strongly elliptic with respect to $`๐๐_+`$ if: (I) The inequalities (30) and (34) hold; (II) There exists a unique solution of the equation
$$[\sigma _L((P+G);\left\{\widehat{x}^k\right\},r=0,\left\{\zeta _j\right\},\omega i\frac{}{r})\lambda ]\phi (r,\widehat{x},\zeta ;\lambda )=0,$$
$`(20^{})`$
subject to the boundary conditions
$$\sigma _L()(\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\})\psi (\phi )=\psi ^{}(\phi )$$
$`(21^{})`$
and to the asymptotic condition (22). It should be stressed that, unlike the case of differential operators, Eq. (20โ) is not an ordinary differential equation in general, because $`(P+G)`$ is pseudo-differential. (III) The strictly homogeneous symbols associated to $`(P+G)`$ and $``$ have limits for $`|\zeta |0`$ in the respective leading symbol norms, with the limiting symbol restricted to the boundary which avoids the values $`\lambda ๐๐_+`$ for all $`\left\{\widehat{x}\right\}`$.
Condition (III) requires a last effort for a proper understanding. Given a pseudo-differential operator of order $`d`$ with leading symbol $`p^0(x,\xi )`$, the associated strictly homogeneous symbol is defined by<sup>12</sup>
$$p^h(x,\xi )|\xi |^dp^0(x,\frac{\xi }{|\xi |})\mathrm{for}\xi 0.$$
$`(37)`$
This extends to a continuous function vanishing at $`\xi =0`$ when $`d>0`$. In the presence of boundaries, the boundary-value problem $`((P+G),)`$ has a strictly homogeneous symbol on the boundary equal to (some indices are omitted for simplicity)
$$\left(\begin{array}{c}p^h(\left\{\widehat{x}\right\},r=0,\left\{\zeta \right\},i\frac{}{r})+g^h(\left\{\widehat{x}\right\},\left\{\zeta \right\},i\frac{}{r})\lambda \\ b^h(\left\{\widehat{x}\right\},\left\{\zeta \right\},i\frac{}{r})\end{array}\right),$$
where $`p^h,g^h`$ and $`b^h`$ are the strictly homogeneous symbols of $`P,G`$ and $``$ respectively, obtained from the corresponding leading symbols $`p^0,g^0`$ and $`b^0`$ via equations analogous to (37), after taking into account the split (19), and upon replacing $`\omega `$ by $`i\frac{}{r}`$. The limiting symbol restricted to the boundary (also called limiting $`\lambda `$-dependent boundary symbol operator) and mentioned in condition III reads therefore<sup>12</sup>
$$\begin{array}{ccc}& a^h(\left\{\widehat{x}\right\},r=0,\zeta =0,i\frac{}{r})\hfill & \\ & =\left(\begin{array}{c}p^h(\left\{\widehat{x}\right\},r=0,\zeta =0,i\frac{}{r})+g^h(\left\{\widehat{x}\right\},\zeta =0,i\frac{}{r})\lambda \\ b^h(\left\{\widehat{x}\right\},\zeta =0,i\frac{}{r})\end{array}\right),\hfill & (38)\hfill \end{array}$$
where the singularity at $`\xi =0`$ of the leading symbol in absence of boundaries is replaced by the singularity at $`\zeta =0`$ of the leading symbols of $`P,G`$ and $``$ when a boundary occurs.
Let us now see how the previous conditions on the leading symbol of $`(P+G)`$ and on the graded leading symbol of the boundary operator can be used. The equation (20โ) is solved by a function $`\phi `$ depending on $`r,\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\}`$ and, parametrically, on the eigenvalues $`\lambda `$. For simplicity, we write $`\phi =\phi (r,\widehat{x},\zeta ;\lambda )`$, omitting indices. Since the leading symbol is no longer a polynomial when $`(P+G)`$ is genuinely pseudo-differential, we cannot make any further specification on $`\phi `$ at this stage, apart from requiring that it should reduce to (here $`|\zeta |^2\zeta _i\zeta ^i`$)
$$\chi (\widehat{x},\zeta )e^{r\sqrt{|\zeta |^2\lambda }}$$
when $`(P+G)`$ reduces to a Laplacian.
The equation (21โ) involves the graded leading symbol of $``$ and restriction to the boundary of the field and its covariant derivative along the normal direction. Such a restriction is obtained by setting to zero the geodesic distance $`r`$, and hence we write, in general form (here we denote again by $`\mathrm{\Lambda }`$ the full matrix element $`_{21}`$ in the boundary operator (10)),
$$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \sigma _L(\mathrm{\Lambda })& I\mathrm{\Pi }\end{array}\right)\left(\begin{array}{c}\phi (0,\widehat{x},\zeta ;\lambda )\\ \phi ^{}(0,\widehat{x},\zeta ;\lambda )\end{array}\right)=\left(\begin{array}{c}\mathrm{\Pi }\rho (0,\widehat{x},\zeta ;\lambda )\\ (I\mathrm{\Pi })\rho ^{}(0,\widehat{x},\zeta ;\lambda )\end{array}\right),$$
$`(39)`$
where $`\rho `$ differs from $`\phi `$, because Eq. (21โ) is written for $`\psi (\phi )`$ and $`\psi ^{}(\phi )\psi (\phi )`$. Now Eq. (39) leads to
$$\mathrm{\Pi }\phi (0,\widehat{x},\zeta ;\lambda )=\mathrm{\Pi }\rho (0,\widehat{x},\zeta ;\lambda ),$$
$`(40)`$
$$\sigma _L(\mathrm{\Lambda })\phi (0,\widehat{x},\zeta ;\lambda )+(I\mathrm{\Pi })\phi ^{}(0,\widehat{x},\zeta ;\lambda )=(I\mathrm{\Pi })\rho ^{}(0,\widehat{x},\zeta ;\lambda ),$$
$`(41)`$
and we require that, for $`\phi `$ satisfying Eq. (20โ) and the asymptotic decay (22), with $`\lambda ๐๐_+`$, Eqs. (40) and (41) can be always solved with given values of $`\rho (0,\widehat{x},\zeta ;\lambda )`$ and $`\rho ^{}(0,\widehat{x},\zeta ;\lambda )`$, whenever $`(\zeta ,\lambda )(0,0)`$. The idea is now to relate, if possible, $`\phi ^{}(0,\widehat{x},\zeta ;\lambda )`$ to $`\phi (0,\widehat{x},\zeta ;\lambda )`$ in such a way that Eq. (40) can be used to simplify Eq. (41). For this purpose, we consider the function $`f`$ such that
$$\frac{\phi ^{}(0,\widehat{x},\zeta ;\lambda )}{\phi (0,\widehat{x},\zeta ;\lambda )}=\frac{\rho ^{}(0,\widehat{x},\zeta ;\lambda )}{\rho (0,\widehat{x},\zeta ;\lambda )}=f(\widehat{x},\zeta ;\lambda ),$$
$`(42)`$
$$\mathrm{\Pi }(\widehat{x})f(\widehat{x},\zeta ;\lambda )=f(\widehat{x},\zeta ;\lambda )\mathrm{\Pi }(\widehat{x}).$$
$`(43)`$
If both (42) and (43) hold, Eq. (41) reduces indeed to
$$\begin{array}{ccc}& \sigma _L(\mathrm{\Lambda })\phi (0,\widehat{x},\zeta ;\lambda )+f(\widehat{x},\zeta ;\lambda )(\phi (0,\widehat{x},\zeta ;\lambda )\rho (0,\widehat{x},\zeta ;\lambda ))\hfill & \\ & =f(\widehat{x},\zeta ;\lambda )\mathrm{\Pi }(\phi (0,\widehat{x},\zeta ;\lambda )\rho (0,\widehat{x},\zeta ;\lambda )),\hfill & (44a)\hfill \end{array}$$
and hence, by virtue of (40),
$$[\sigma _L(\mathrm{\Lambda })+f(\widehat{x},\zeta ;\lambda )]\phi (0,\widehat{x},\zeta ;\lambda )=\rho ^{}(0,\widehat{x},\zeta ;\lambda ).$$
$`(44b)`$
Thus, the strong ellipticity condition with respect to $`๐๐_+`$ implies in this case the invertibility of $`[\sigma _L(\mathrm{\Lambda })+f(\widehat{x},\zeta ;\lambda )]`$, i.e.
$$\mathrm{det}[\sigma _L(\mathrm{\Lambda })+f(\widehat{x},\zeta ;\lambda )]0\lambda ๐๐_+.$$
$`(45)`$
Moreover, by virtue of the identity
$$[f(\widehat{x},\zeta ;\lambda )+\sigma _L(\mathrm{\Lambda })][f(\widehat{x},\zeta ;\lambda )\sigma _L(\mathrm{\Lambda })]=[f^2(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })],$$
$`(46)`$
the condition (45) is equivalent to
$$\mathrm{det}[f^2(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })]0\lambda ๐๐_+.$$
$`(47)`$
Since $`f(\widehat{x},\zeta ;\lambda )`$ is, in general, complex-valued, one can always express it in the form
$$f(\widehat{x},\zeta ;\lambda )=\mathrm{Re}f(\widehat{x},\zeta ;\lambda )+i\mathrm{Im}f(\widehat{x},\zeta ;\lambda ),$$
$`(48)`$
so that (47) reads eventually
$$\mathrm{det}[\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\mathrm{Im}^2f(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })+2i\mathrm{Re}f(\widehat{x},\zeta ;\lambda )\mathrm{Im}f(\widehat{x},\zeta ;\lambda )]0.$$
$`(49)`$
In particular, when
$$\mathrm{Im}f(\widehat{x},\zeta ;\lambda )=0,$$
$`(50)`$
condition (49) reduces to
$$\mathrm{det}[\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })]0.$$
$`(51)`$
A sufficient condition for strong ellipticity with respect to the cone $`๐๐_+`$ is therefore the negative-definiteness of $`\sigma _L^2(\mathrm{\Lambda })`$:
$$\sigma _L^2(\mathrm{\Lambda })<0,$$
$`(52)`$
so that
$$\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })>0,$$
$`(53)`$
and hence (51) is fulfilled.
In the derivation of the sufficient conditions (49) and (52), the assumption (43) plays a crucial role. In general, however, $`\mathrm{\Pi }`$ and $`f`$ have a non-vanishing commutator, and hence a $`C(\widehat{x},\zeta ;\lambda )`$ exists such that
$$\mathrm{\Pi }(\widehat{x})f(\widehat{x},\zeta ;\lambda )f(\widehat{x},\zeta ;\lambda )\mathrm{\Pi }(\widehat{x})=C(\widehat{x},\zeta ;\lambda ).$$
$`(54)`$
The occurrence of $`C`$ is a peculiar feature of the fully pseudo-differential framework. Equation (41) is then equivalent to (now we write explicitly also the independent variables in the leading symbol of $`\mathrm{\Lambda }`$)
$$\begin{array}{ccc}& [(\sigma _L(\mathrm{\Lambda })C)(\widehat{x},\zeta ;\lambda )+f(\widehat{x},\zeta ;\lambda )]\phi (0,\widehat{x},\zeta ;\lambda )\hfill & \\ & =\rho ^{}(0,\widehat{x},\zeta ;\lambda )C(\widehat{x},\zeta ;\lambda )\rho (0,\widehat{x},\zeta ;\lambda ).\hfill & (55)\hfill \end{array}$$
On defining
$$\gamma (\widehat{x},\zeta ;\lambda )[\sigma _L(\mathrm{\Lambda })C](\widehat{x},\zeta ;\lambda ),$$
$`(56)`$
we therefore obtain strong ellipticity conditions formally analogous to (45) or (49) or (51), with $`\sigma _L(\mathrm{\Lambda })`$ replaced by $`\gamma (\widehat{x},\zeta ;\lambda )`$ therein, i.e.
$$\mathrm{det}[\gamma (\widehat{x},\zeta ;\lambda )+f(\widehat{x},\zeta ;\lambda )]0\lambda ๐๐_+,$$
$`(57)`$
which is satisfied if
$$\mathrm{det}[\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\mathrm{Im}^2f(\widehat{x},\zeta ;\lambda )\gamma ^2(\widehat{x},\zeta ;\lambda )+2i\mathrm{Re}f(\widehat{x},\zeta ;\lambda )\mathrm{Im}f(\widehat{x},\zeta ;\lambda )]0.$$
$`(58)`$
We have therefore provided a complete characterization of the properties of the symbol of the boundary operator for which a set of boundary conditions completely invariant under infinitesimal diffeomorphisms are compatible with a strongly elliptic one-loop quantum theory. Our analysis is detailed but general, and hence has the merit (as far as we can see) of including all pseudo-differential boundary operators for which the sufficient conditions just derived can be imposed. This is not yet the same, however, as saying that the pseudo-differential framework in one-loop quantum theory is definitely better. One still has to prove that the set of symbols satisfying all our conditions is non-empty. Moreover, our definition of strong ellipticity is given for self-adjoint pseudo-differential boundary-value problems, and is therefore less general than the one applied in Ref. 7.
It would be now very interesting to prove that, by virtue of the pseudo-differential nature of $``$ in (10), the quantum state of the universe in one-loop semiclassical theory can be made of surface-state type.<sup>14</sup> This would describe a wave function of the universe with exponential decay away from the boundary, which might provide a novel description of quantum physics at the Planck length. It therefore seems that by insisting on path-integral quantization, strong ellipticity of the Euclidean theory and invariance principles, new deep perspectives are in sight. These are in turn closer to what we may hope to test, i.e. the one-loop semiclassical approximation in quantum gravity. In the seventies, such calculations could provide a guiding principle for selecting couplings of matter fields to gravity in a unified field theory. Now they can lead instead to a deeper understanding of the interplay between non-local formulations,<sup>15-17</sup> elliptic theory, gauge-invariant quantization<sup>18</sup> and a quantum theory of the very early universe.<sup>10</sup>
Acknowledgments
The author is indebted to the Organizers for giving him the opportunity to submit this invited contribution, to Ivan Avramidi for scientific collaboration and to Gerd Grubb for detailed correspondence.
References
1. B. S. DeWitt, Dynamical Theory of Groups and Fields (Gordon and Breach, New York, 1965).
2. B. S. DeWitt, in Relativity, Groups and Topology II, eds. B. S. DeWitt and R. Stora (North-Holland, Amsterdam, 1984).
3. R. P. Feynman, Rev. Mod. Phys. 20, 367 (1948).
4. C. W. Misner, Rev. Mod. Phys. 29, 497 (1957).
5. S. W. Hawking, in General Relativity, an Einstein Centenary Survey, eds. S. W. Hawking and W. Israel (Cambridge University Press, Cambridge, 1979).
6. I. G. Avramidi and G. Esposito, Commun. Math. Phys. 200, 495 (1999).
7. G. Grubb, Ann. Scuola Normale Superiore Pisa 1, 1 (1974).
8. P. B. Gilkey and L. Smith, J. Diff. Geom. 18, 393 (1983).
9. G. Esposito, Class. Quantum Grav. 16, 3999 (1999).
10. G. Esposito, โBoundary Operators in Quantum Field Theoryโ (HEP-TH 0001086, to appear in Int. J. Mod. Phys. A).
11. P. B. Gilkey, Invariance Theory, the Heat Equation and the AtiyahโSinger Index theorem (Chemical Rubber Company, Boca Raton, 1995).
12. G. Grubb, Functional Calculus of Pseudodifferential Boundary Problems (Birkhรคuser, Boston, 1996).
13. G. Esposito, Class. Quantum Grav. 16, 1113 (1999).
14. M. Schrรถder, Rep. Math. Phys. 27, 259 (1989).
15. J. W. Moffat, Phys. Rev. D41, 1177 (1990).
16. D. Evens, J. W. Moffat, G. Kleppe and R. P. Woodard, Phys. Rev. D43, 499 (1991).
17. V. N. Marachevsky and D. V. Vassilevich, Class. Quantum Grav. 13, 645 (1996).
18. G. Esposito and C. Stornaiolo, Int. J. Mod. Phys. A15, 449 (2000).
|
warning/0006/cond-mat0006159.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Strongly nonequilibrium cooperative processes that occur in statistical systems interacting with electromagnetic field are described by complicated nonlinear differential and integroโdifferential equations. For treating such difficult problems, a general approach has been recently developed called the Scale Separation Approach whose basic idea is to present the evolution equations in such a form where it could be possible to separate several characteristic spaceโtime scales. In many cases, different scales appear rather naturally being directly related to the physical properties of the considered system.
Since the scale separation approach makes the mathematical foundation for the following applications, we start the review with presenting the basic techniques of this approach. Then we demonstrate it by applying the method to different physical problems related to strongly nonequilibrium processes occurring under the interaction of electromagnetic field with matter. The considered examples not only serve as illustrations of the method but are of importance as such since they concern interesting and rather nontrivial physical effects. For consideration, those effects are chosen that have been first correctly described or predicted by the authors. Among these effects, we would like to emphasize, as the most interesting, the following: Collective Liberation of Light, Turbulent Photon Filamentation, Superradiant Spin Relaxation, Negative Electric Current, and Magnetic Semiconfinement of Atoms.
The content of the report is as follows. In Section 2 the Scale Separation Approach is described. This method makes it possible to solve, or to strongly simplify, many complicated systems of nonlinear differential equations, including stochastic and partialโderivative equations. The mathematical procedure of solving nonlinear differential equations in the following applications is based on this approach. The examples we consider have mainly to do with the evolution equations describing strongly nonequilibrium statistical systems interacting with electromagnetic fields. We concentrate our attention on collective phenomena whose existence as such, as well as their properties, are due to nonlinear effects. This is why we constantly have to deal with nonlinear equations.
Resonant interactions of electromagnetic field with radiating systems are usually describing by the MaxwellโBloch equations, in which one often passes to the momentum representation by means of Fourier transform. But we prefer to work in the RealโSpace representation, outlined in Section 3, which seems to be more convenient for employing the Scale Separation Approach. Another convenient trick we employ is the elimination of electromagnetic field from evolution equations. For this purpose, the operator Maxwell equations, supplemented by the Coulomb calibration, can be rewritten in the integral form connecting the vector potential with the retarded current formed by the radiating system. Substituting this vector potential into evolution equations eliminates from them electromagnetic field. In this way, we come to the system of equations not containing explicitly electromagnetic field, instead of which there appears an effective dipole interaction of radiating atoms. After eliminating electromagnetic field, we have less equations, although the price for this is that these equations become integroโdifferential. nevertheless, the obtained equations are more convenient for applying to them our method of solution. Another important advantage of the derived equations is the possibility of taking into account quantum effects. Such effects are often principal, while the standard semiclassical MaxwellโBloch equations cannot take account of them. To simplify evolution equations, not loosing quantum effects, is the idea of the Stochastic MeanโField Approximation of Section 4. Since cooperative electromagnetic phenomena are directly related to arising coherence, Section 5 gives the definitions for Dynamical Characteristics of Coherence.
The equations derived in the previous sections and the method of solution developed above are applied to several concrete systems exhibiting interesting physical properties. In Section 6, we suggest the theory of Cอกollective Liberation of Light, which can occur in materials with polariton band gap. In Section 7, we consider the influence of external fields on radiation properties of resonant atoms, checking whether it is feasible to get Amplification by Nonresonant Fields. Section 8 discusses the soโcalled Mรถssbauer Magnetic Anomaly observed in some magnetic materials. In Section 9 the Problem of Pattern Selection is analyzed. This problem arises, for instance, when one needs to describe resonant media with spatially nonuniform electromagnetic structures. For treating the problem, we have suggested an original approach based on probabilistic analysis of possible spatioโtemporal patterns. This method is applied, in Section 10, to describing Turbulent Photon Filamentation in resonant media.
Scale Separation Approach, being a general method, can be employed for treating strongly nonequilibrium systems of different physical nature. In Section 11, it is used for giving a thorough picture of Superradiant Spin Relaxation occurring in nonequilibrium nuclear magnets. This method also makes it possible to analyse nonlinear differential equations in partial derivatives. Such an analysis helps in finding conditions under which unusual nonlinear effects can happen. This is illustrated in Section 12 by describing a transient effect of Negative Electric Current in nonuniform semiconductors. Another novel effect of Magnetic Semiconfinement of Atoms is described in Section 13. Both these effects have been predicted by the authors. In Section 14, we discuss conditions when Nuclear Matter Lasing could be possible.
Throughout the review, we consider several physical systems of rather different nature, Because of this, it is more appropriate to give all details and to discuss the related literature in the corresponding sections, limiting the Introduction by a brief enumeration of the considered problems. Section 15 contains Conclusion summarizing main results.
## 2 Scale Separation Approach
Because of the pivotal role of this approach for treating physical problems in the following sections, we need to start by presenting its general scheme. It is possible to separate five main steps, or parts, of the approach: (i) stochastic quantization of shortโrange correlations; (ii) separation of variables onto fast and slow; (iii) averaging method for multifrequency systems; (iv) generalized expansion about guiding centers; and (v) selection of scales for space structures. Below, these steps are explicitly explained.
2.1. ShortโRange Stochastic Quantization. When considering nonequilibrium processes in statistical systems, one needs to write evolution equations for some averages $`<A_i>`$ of operators $`A_i(t)`$, where $`t`$ is time and $`i=1,2,\mathrm{},N`$ enumerates particles composing the considered system. For simplicity, a discrete index $`i`$ is employed, although everywhere in what follows one could mean an operator $`A(\stackrel{}{r}_i,t)`$ depending on a continuous space variable $`\stackrel{}{r}_i`$.
There exists the wellโknown problem in statistical mechanics consisting in the fact that writing an evolution equation for $`<A_i>`$ one does not get a closed set of equations but a hierarchical chain of equations connecting correlation functions of higher orders. Thus, an equation for $`<A_i>`$ involves the terms as $`_j<A_iB_j>`$ with double correlators $`<A_iB_j>`$, and the evolution equations for the latter acquire the terms with triple correlators, and so on. The simplest way for making the system of equations closed is by resorting to the meanโfield type decoupling $`<A_iB_j><A_i><B_j>`$. When considering radiation processes, this decoupling is called the semiclassical approximation. Then the term $`_j<A_iB_j>`$ reduces to $`<A_i>_j<B_j>`$, so that one can say that $`<A_i>`$ is subject to the action of the mean field $`_j<B_j>`$. The semiclassical approximation describes well coherent processes, when longโrange correlations between particles govern the evolution of the system, while shortโrange correlations, due to quantum fluctuations, are not important. However, the latter may become of great importance if there are periods of time when the longโrange correlations are absent. For example, this may happen at the beginning of a nonequilibrium process when longโrange correlations have had yet no time to develop. Then neglecting shortโrange correlations can lead to principally wrong results for the whole dynamics.
To include the influence of shortโrange correlations, the semiclassical approximation can be modified as follows:
$$\underset{j}{}<A_iB_j>=<A_i>\left(\underset{j}{}<B_j>+\xi \right),$$
(1)
where $`\xi `$ is a random variable describing local shortโrange correlations. It is natural to treat $`\xi `$ as a Gaussian stochastic variable defined by its first, $`\xi `$, and second, $`|\xi |^2`$, moments. According to the shortโrange character of local fields, we should set
$$\xi =\mathrm{\hspace{0.33em}0}.$$
(2)
The second moment, aiming at taking into account incoherent local fluctuations, can be defined by means of the following reasoning. Consider the equality
$$|\underset{j}{}<A_iB_j>|^2=|<A_i>|^2\left(\right|\underset{j}{}<B_j>|^2+|\xi |^2)$$
resulting from definitions (1) and (2). On the other hand, wishing to take into account both longโrange coherent as well as shortโrange incoherent terms, one should write
$$\left|\underset{j}{}<A_iB_j>\right|^2=|<A_i>|^2\left(\left|\underset{j}{}<B_j>\right|^2+\underset{j}{}|<B_j>|^2\right),$$
where the first term in the brackets corresponds to the coherent while the second term, to incoherent parts. Comparing the latter two equalities, we come to the conclusion that
$$|\xi |^2=\underset{j}{}|<B_j>|^2.$$
(3)
As far as shortโrange correlations and fluctuations are often due to quantum effects, the manner of taking them into account by introducing a stochastic variable $`\xi `$ can be named the stochastic quantization. Then the decoupling (1) may be termed the stochastic meanโfield approximation. A similar kind of approximation has been used for taking account of quantum spontaneous emission of atoms in the problem of atomic superradiance . Somewhat related ideas have also been used in the stochastic quantization of quantum field theory .
2.2. Classification of Function Variations. Employing the stochastic meanโfield approximation makes it possible to write down a closed set of stochastic differential equations. The next step is to find such a change of variables which results in the possibility of separating the functional variables onto fast and slow. Let us consider, first, the variation of functions in time. Assume that we come to the set of equations of the form
$$\frac{du}{dt}=f,\frac{ds}{dt}=\epsilon g,$$
(4)
in which $`f=f(\epsilon ,u,s,\xi ,t),g=g(\epsilon ,u,s,\xi ,t)`$, and $`\epsilon 1`$ is a small parameter. Equations (4) are complimented by initial conditions
$$u=u_0,s=s_0(t=0).$$
(5)
Here, for simplicity, we deal with only two functions, $`u`$ and $`s`$, and one small parameter $`\epsilon `$. The whole procedure is straightforwardly applicable to the case of many functions and several parameters.
Let the functions $`f`$ and $`g`$ be such that
$$\underset{\epsilon 0}{lim}f0,\underset{\epsilon 0}{lim}\epsilon g=0.$$
(6)
Then from Eqs. (4) it follows that
$$\underset{\epsilon 0}{lim}\frac{du}{dt}0,\underset{\epsilon 0}{lim}\frac{ds}{dt}=0.$$
(7)
This permits us to classify the solution $`u`$ as fast, compared to the slow solution $`s`$. In turn, the slow solution $`s`$ is a quasiโinvariant with respect to the fast solution $`u`$. Thus, we may classify the functions representing the sought solutions onto fastly and slowly varying in time.
In the case of partial differential equations, one has, in addition to time, a space variable $`\stackrel{}{r}`$. Then the notion of fast and slow functions can be generalized as follows . Let $`\stackrel{}{r}๐`$, with $`V`$ being the measure of the volume $`๐`$, and let $`t[0,T]`$, where $`T`$ can be infinite. If one has
$$\underset{\epsilon 0}{lim}\frac{1}{V}_๐\frac{u}{t}d\stackrel{}{r}0,\underset{\epsilon 0}{lim}\frac{1}{T}_0^T\stackrel{}{}udt0,$$
(8)
while
$$\underset{\epsilon 0}{lim}\frac{1}{V}_๐\frac{s}{t}d\stackrel{}{r}=0,\underset{\epsilon 0}{lim}\frac{1}{T}_0^T\stackrel{}{}sdt=0,$$
(9)
then the solution $`u`$ can be called fast on average with respect to both space and time, as compared to $`s`$ that is slow on average. In such a case, $`s`$ is again a quasiโinvariant with respect to $`u`$. In general, it may, of course, happen that one of the solutions is fast in time but slow in space, or vice versa, as compared to another solution. Note that in the Hamiltonian mechanics quasiโinvariants with respect to time are called adiabatic invariants . A generalization of this notion to the case of both space and time variables is given by definition (9).
2.3. Multifrequency Averaging Technique. Let us continue considering the ordinary differential equations (4). The generalization to the case of partial differential equations can be done similarly to the way discussed at the end of the previous section. After classifying the function $`u`$ as fast and $`s`$ as slow, we may resort to the KrylovโBogolubov averaging technique extended to multifrequency systems.
Since the slow solution $`s`$ is a quasiโinvariant for the fast variable $`u`$, one considers the equation for the fast function, with the slow one kept fixed,
$$\frac{X}{t}=f(\epsilon ,X,z,\xi ,t),$$
(10)
here $`s=z`$ being treated as a constant parameter. The initial conditions for Eq. (10) is
$$X=u_0(t=0).$$
(11)
The pair of solutions
$$X=X(\epsilon ,z,\xi ,t),z=const$$
(12)
are called the generating solutions. Substituting the solution $`X`$ into the rightโhand side of the equation for the slow function $`s`$, one defines the average
$$\overline{g}(\epsilon ,z)\frac{1}{\tau }_0^\tau g(\epsilon ,X(\epsilon ,z,\xi ,t),z,\xi ,t)dt,$$
(13)
in which $`\tau `$ is the characteristic time of fast oscillations. In many cases, it is sufficient to set $`\tau \mathrm{}`$. In this way, we come to the equation
$$\frac{dz}{dt}=\epsilon \overline{g}(\epsilon ,z),$$
(14)
with the initial condition
$$z=s_0(t=0).$$
(15)
The solution to Eq. (14),
$$z=z(\epsilon ,t),$$
(16)
is to be substituted into $`X`$ yielding
$$y(\epsilon ,\xi ,t)=X(\epsilon ,z(\epsilon ,t),\xi ,t).$$
(17)
Generating solutions (12) are the first crude approximations one starts with. More elaborate solutions (16) and (17) are termed guiding centers.
Notice two points that difference the considered way of obtaining the guiding centers (16) and (17) from the standard averaging method . The first point is in retaining in Eq. (10) the small parameter $`\epsilon `$, which makes it possible to correctly take into account important physical effects, such as attenuation. The standard manner of defining the generating solutions with setting $`\epsilon =0`$ would result in essentially more rough approximations. The second difference is in the occurrence of the stochastic average in definition (13), since here we are dealing with stochastic differential equations.
2.4. Generalized Asymptotic Expansion. The generating solutions (12) play the role of the trial zeroโorder approximation, while the guiding centers (16) and (17) essentially improve the trial approximations. Higherโorder corrections may be obtained by presenting the general solutions as asymptotic expansions about the guiding centers. Then, $`k`$โorder approximations are written as
$$u_k=y(\epsilon ,\xi ,t)+\underset{n=1}{\overset{k}{}}y_n(\epsilon ,\xi ,t)\epsilon ^n,s_k=z(\epsilon ,t)+\underset{n=1}{\overset{k}{}}z_n(\epsilon ,\xi ,t)\epsilon ^n.$$
(18)
Such series are named generalized asymptotic expansions , since the expansion coefficients depend themselves on parameter $`\epsilon `$. The rightโhand sides of Eqs. (4) are also to be expanded about the guiding centers yielding
$$f(\epsilon ,u_k,s_k,\xi ,t)f(\epsilon ,y,z,\xi ,t)+\underset{n=1}{\overset{k}{}}f_n(\epsilon ,\xi ,t)\epsilon ^n,$$
$$g(\epsilon ,u_k,s_k,\xi ,t)g(\epsilon ,y,z,\xi ,t)+\underset{n=1}{\overset{k}{}}g_n(\epsilon ,\xi ,t)\epsilon ^n.$$
(19)
Then, expansions (18) and (19) are to be substituted in Eqs. (4) with equating the like terms with respect to the explicit powers of $`\epsilon `$. Thus, in the first order, this gives
$$\frac{dy_1}{dt}=f_1(\epsilon ,\xi ,t)\overline{g}(\epsilon ,z)X_1(\epsilon ,\xi ,t),\frac{dz_1}{dt}=g(\epsilon ,y,z,\xi ,t)\overline{g}(\epsilon ,z),$$
(20)
where
$$X_1(\epsilon ,\xi ,t)\frac{}{z}X(\epsilon ,z,\xi ,t),z=z(\epsilon ,t).$$
For the approximations of order $`n2`$, we get
$$\frac{dy_n}{dt}=f_n(\epsilon ,\xi ,t),\frac{dz_n}{dt}=g_n(\epsilon ,\xi ,t).$$
(21)
The initial conditions for all $`n=1,2,\mathrm{}`$ are
$$y_n=z_n=0(t=0).$$
(22)
The functions $`f_n`$ and $`g_n`$ depend on $`y_1,y_2,\mathrm{},y_n`$, and on $`z_1,z_2,\mathrm{},z_n`$, but it is important that the dependence on $`y_n`$ and $`z_n`$ is linear. The latter follows from the fact that expanding a function
$$f\left(y+\underset{n=1}{\overset{k}{}}y_n\epsilon ^n\right)=\underset{n=1}{\overset{k}{}}f_n\epsilon ^n$$
in powers of $`\epsilon `$, one has
$$f_1=f^{}(y)y_1,f_2=\frac{1}{2!}\left[f^{\prime \prime }(y)y_1+f^{}(y)y_2\right],f_3=\frac{1}{3!}\left[f^{\prime \prime \prime }(y)y_1+2f^{\prime \prime }(y)y_2+f^{}(y)y_3\right],$$
and so on. In this way, Eqs. (20) directly define $`y_1`$ and $`z_1`$, and Eqs. (21) are linear equations, thus, being easily integrated.
Usually, one does not need the higherโorder approximations since the main physics, in the majority of cases, is already well described by the guiding centers (16) and (17). The latter are good approximations to the exact solutions in the time interval $`0tT_s/\epsilon `$, where $`T_s`$ is a characteristic time of the slowโsolution variation. In those cases when the higherโorder approximations are important, each $`k`$โorder approximant can also be improved by invoking some sort of summation of asymptotic series (18), for instance, the selfโsimilar summation \[9โ12\].
2.5. Selection of Space Structures. The solutions to differential or integroโdifferential nonlinear equations in partial derivatives are generally nonuniform in space exhibiting the formation of different spatial structures. And it often happens that a given set of equations possesses several solutions corresponding to different spatial patterns or to different scales of such patterns . When there is a family of solutions describing several possible patterns, the question arises which of these solutions, and respectively patterns, is preferable and in what sense could it be preferable. This problem of pattern selection is a general and very important problem constantly arising in considering spatial structures. In this subsection we delineate a simple way that in many cases helps to solve the problem of pattern selection. A more refined theory will be presented in sections 9 and 10.
Assume that the obtained solutions describe spatial structures that can be parametrized by a multiparameter $`\beta `$, so that the $`k`$โorder approximations $`u_k(\beta ,\stackrel{}{r},t)`$ and $`s_k(\beta ,\stackrel{}{r},t)`$ include the dependence on $`\beta `$ whose value is, however, yet undefined. To define $`\beta `$, and respectively the related pattern, one may proceed in the spirit of the selfโsimilar approximation theory \[14โ23\], by treating $`\beta `$ as a control function. According to the theory \[14โ23\], control functions are to be defined from fixedโpoint conditions for an approximation cascade constructed for an observable quantity. For the latter, one may take the average energy defined as follows. The internal energy, which is a statistical average of the system Hamiltonian, is a functional $`E[u,s]`$ of the solutions. Taking the $`k`$โorder approximations for the latter and averaging the internal energy over the period of fast oscillations and over stochastic variables, one gets the average energy
$$E_k(\beta )\frac{1}{\tau }_0^\tau E[u_k(\beta ,\stackrel{}{r},t),s_k(\beta ,\stackrel{}{r},t)]dt.$$
(23)
For the sequence of approximations, $`\{E_k(\beta )\}`$, it is possible to construct an approximation cascade whose fixed point can be given by the condition
$$\frac{}{\beta }E_k(\beta )=0,$$
(24)
from which one gets the control function $`\beta =\beta _k`$ defining the corresponding pattern. According to optimal control theory, control functions are defined so that to minimize a cost functional. The latter, in our case, is naturally represented by the average energy (23). Hence, when the fixedโpoint equation (24) has several solutions, one may select of them that one which minimizes the cost functional (23), so that
$$E_k(\beta _k)=\mathrm{abs}\underset{\beta }{\mathrm{min}}E_k(\beta ).$$
(25)
Equations (24) and (25) have a simple physical interpretation as the minimum conditions for the average energy (23). However, one should keep in mind that there is no, in general, such a principle of minimal energy for nonequilibrium systems . Therefore the usage of the ideas from the selfโsimilar approximation theory \[14โ23\] provides a justification for employing conditions (24) and (25) for nonequilibrium processes.
The scale separation approach presented in this section makes it possible to solve rather complicated sets of nonlinear differential equations describing various nonequilibrium phenomena in statistical systems. More details on this approach can be found in Refs. \[24โ28\].
## 3 Real Space Representation
When considering the interaction of atoms with electromagnetic fields, one usually employs the soโcalled mode representation, expanding field operators over mode wave functions . These can be either freeโmode functions, that is plane waves, or resonatorโmode functions depending on the resonator geometry. We prefer to deal with the realโspace representation because of the following reasons: First, the evolution equations in this representation are written in a form more convenient for analysing temporal nonstationary behaviour of solutions. Second, it is more suitable for describing nonuniform solutions corresponding to selfโorganized space structures. And third, this representation is more appropriate for using the scale separation approach. Since the real space representation is rarely considered in literature, it is worth recalling in brief the derivation of the main equations in this representation . To understand the basis of the main evolution equations is very important, for these equations will be constantly used in what follows. One more peculiarity of the consideration below, differencing it form the standard texts, is the comparison of the formulas for the cases of electrodipole and magnitodipole transitions.
Let us have a system of radiators that can be atoms, molecules, nuclei, etc. Assume that the size of a radiator, $`a_0`$, is small as compared to the mean distance between them, $`a`$, as well as to the characteristic radiation wavelength $`\lambda `$,
$$\frac{a_0}{a}1,\frac{a_0}{\lambda }1,$$
(26)
while the relation between $`a`$ and $`\lambda `$ can be arbitrary. Canonical variables related to the electromagnetic field are the electric field $`\stackrel{}{E}`$ and the vector potential $`\stackrel{}{A}`$, whose commutation relations are
$$[E^\alpha (\stackrel{}{r},t),A^\beta (\stackrel{}{r}^{},t)]=4\pi ic\delta _{\alpha \beta }\delta (\stackrel{}{r}\stackrel{}{r}^{}),$$
$$[E^\alpha (\stackrel{}{r},t),E^\beta (\stackrel{}{r}^{},t)]=[A^\alpha (\stackrel{}{r},t),A^\beta (\stackrel{}{r}^{},t)]=0,$$
where $`c`$ is the light velocity and the index $`\alpha ,\beta =1,2,3`$, or $`x,y,z`$, enumerate the Cartesian coordinates. The magnetic field is
$$\stackrel{}{H}(\stackrel{}{r},t)=\stackrel{}{}\times \stackrel{}{A}(\stackrel{}{r},t).$$
To uniquely define the latter, we invoke the Coulomb gauge condition
$$\stackrel{}{}\stackrel{}{A}(\stackrel{}{r},t)=0.$$
Here and in what follows the system of units is used where $`\mathrm{}1`$.
The radiator charges are described by the annihilation, $`\psi `$, and creation, $`\psi ^{}`$, field operators with the commutation relations
$$[\psi (\stackrel{}{r},t),\psi ^{}(\stackrel{}{r}^{},t)]_{}=\delta (\stackrel{}{r}\stackrel{}{r}^{}),[\psi (\stackrel{}{r},t),\psi (\stackrel{}{r}^{},t)]_{}=0,$$
$$[\psi (\stackrel{}{r},t),\stackrel{}{E}(\stackrel{}{r}^{},t)]=[\psi (\stackrel{}{r},t),\stackrel{}{A}(\stackrel{}{r}^{},t)]=0,$$
in which the indices minus or plus mean the commutators or anticommutators, respectively, depending on the Bose or Fermi statistics of the charges.
Assume that in addition to the quantum radiation fields $`\stackrel{}{E}`$ and $`\stackrel{}{H}`$ there are classical fields $`\stackrel{}{E}_0`$ and $`\stackrel{}{H}_0`$ for which we have
$$\stackrel{}{E}_0(\stackrel{}{r},t)=\stackrel{}{}\phi _0(\stackrel{}{r},t),\stackrel{}{H}_0(\stackrel{}{r},t)=\stackrel{}{}\times \stackrel{}{A}_0(\stackrel{}{r},t),\stackrel{}{}\stackrel{}{A}_0(\stackrel{}{r},t)=0.$$
These additional fields can be due to external sources or can be created by the matter which the radiators are inserted in.
Each radiator is also subject to the action of a scalar potential $`\phi _i(\stackrel{}{r})`$ representing all stationary Coulomb interactions. Thus, we may introduce the total scalar and vector potentials
$$\phi _{tot}(\stackrel{}{r},t)=\phi _0(\stackrel{}{r},t)+\underset{i=1}{\overset{N}{}}\phi _i(\stackrel{}{r}),\stackrel{}{A}_{tot}(\stackrel{}{r},t)=\stackrel{}{A}_0(\stackrel{}{r},t)+\stackrel{}{A}(\stackrel{}{r},t),$$
(27)
where $`N`$ is the number of radiators. Then the local energy operator is defined as
$$\widehat{H}(\stackrel{}{r},t)=\frac{1}{2m_0}\left[i\stackrel{}{}+\frac{e}{c}\stackrel{}{A}_{tot}(\stackrel{}{r},t)\right]^2+e\phi _{tot}(\stackrel{}{r},t),$$
(28)
where $`m_0`$ is mass and $`e`$, charge of a particle. Omitting here the relativistic term $`e^2\stackrel{}{A}_{tot}^2/c^2`$ and using the Coulomb calibration, we have
$$\widehat{H}(\stackrel{}{r},t)=\frac{^2}{2m_0}+\frac{ie}{m_0c}\stackrel{}{A}_{tot}(\stackrel{}{r},t)\stackrel{}{}+e\phi _{tot}(\stackrel{}{r},t).$$
The Hamiltonian of the system of radiators interacting with electromagnetic field and with matter is written as the sum
$$\widehat{H}=\widehat{H}_r+\widehat{H}_f+\widehat{H}_{rf}+\widehat{H}_m+\widehat{H}_{mf},$$
(29)
in which the terms represent, respectively, the Hamiltonians of radiators, field, radiatorโfield interaction, matter, and matterโfield interaction. The Hamiltonian of the system of radiators is
$$\widehat{H}_r(t)=\psi ^{}(\stackrel{}{r},t)\left[\frac{^2}{2m_0}+e\underset{i=1}{\overset{N}{}}\phi _i(\stackrel{}{r})\right]\psi (\stackrel{}{r},t)๐\stackrel{}{r}.$$
(30)
This includes also the direct interaction of radiators with matter by means of the effective scalar potentials $`\phi _i(\stackrel{}{r})`$. The field Hamiltonian writes
$$\widehat{H}_f(t)=\frac{1}{8\pi }\left[\stackrel{}{E}^2(\stackrel{}{r},t)+\stackrel{}{H}^2(\stackrel{}{r},t)\right]๐\stackrel{}{r}.$$
(31)
The radiatorโfield interaction is described by
$$\widehat{H}_{rf}(t)=\psi ^{}(\stackrel{}{r},t)\left[\frac{ie}{m_0c}\stackrel{}{A}_{tot}(\stackrel{}{r},t)\stackrel{}{}+e\phi _0(\stackrel{}{r},t)\right]\psi (\stackrel{}{r},t)๐\stackrel{}{r}.$$
(32)
The Hamiltonians of matter and of matterโfield interaction are to be specified according to particular cases under consideration.
The size of a radiator, according to inequalities (26), is the smallest characteristic length. If $`\stackrel{}{r}_i`$ is the centerโofโmass of a radiator, we shall use the notation
$$\stackrel{}{E}_i(t)\stackrel{}{E}(\stackrel{}{r}_i,t),\stackrel{}{H}_i(t)=\stackrel{}{H}(\stackrel{}{r}_i,t),$$
$$\stackrel{}{A}_i(t)\stackrel{}{A}(\stackrel{}{r}_i,t),\stackrel{}{E}_{0i}(t)\stackrel{}{E}_0(\stackrel{}{r}_i,t),\stackrel{}{H}_{0i}(t)=\stackrel{}{H}_0(\stackrel{}{r}_i,t).$$
For $`\stackrel{}{r}`$ in the vicinity of $`\stackrel{}{r}_i`$, we may write
$$\phi _0(\stackrel{}{r},t)\stackrel{}{r}\stackrel{}{E}_{0i}(t)(\stackrel{}{r}\stackrel{}{r}_i),$$
$$\stackrel{}{A}_0(\stackrel{}{r},t)\frac{1}{2}\stackrel{}{r}\times \stackrel{}{H}_{0i}(t),\stackrel{}{A}(\stackrel{}{r},t)\stackrel{}{A}_i(t)\frac{1}{2}(\stackrel{}{r}\stackrel{}{r}_i)\times \stackrel{}{H}_i(t).$$
The energy levels of each radiator are defined by the Schrรถdinger equation
$$\left[\frac{^2}{2m_0}+e\phi _i(\stackrel{}{r})\right]\psi _n(\stackrel{}{r}\stackrel{}{r}_i)=E_n\psi _n(\stackrel{}{r}\stackrel{}{r}_i),$$
where it is assumed that all radiators are identical and $`\phi _i(\stackrel{}{r})=\phi (\stackrel{}{r}\stackrel{}{r}_i)`$. The eigenfunctions $`\psi _n(\stackrel{}{r}\stackrel{}{r}_i)`$ form a complete orthonormal set enumerated by the indices $`n`$ and $`i`$, so that
$$\psi _m^{}(\stackrel{}{r}\stackrel{}{r}_i)\psi _n(\stackrel{}{r}\stackrel{}{r}_j)๐\stackrel{}{r}=\delta _{mn}\delta _{ij},\underset{in}{}\psi _n^{}(\stackrel{}{r}\stackrel{}{r}_i)\psi _n(\stackrel{}{r}^{}\stackrel{}{r}_i)=\delta (\stackrel{}{r}\stackrel{}{r}^{}).$$
With these functions, we may define the density of transition current
$$\stackrel{}{j}_{mn}(\stackrel{}{r})=\frac{ie}{2m_0}\left[\psi _m^{}(\stackrel{}{r})\stackrel{}{}\psi _n(\stackrel{}{r})\psi _n(\stackrel{}{r})\stackrel{}{}\psi _m^{}(\stackrel{}{r})\right]$$
(33)
and the transition current
$$\stackrel{}{j}_{mn}=\stackrel{}{j}_{mn}(\stackrel{}{r})๐\stackrel{}{r}.$$
(34)
We also introduce the electric transition dipole
$$\stackrel{}{d}_{mn}=e\psi _m^{}(\stackrel{}{r})\stackrel{}{r}\psi _n(\stackrel{}{r})๐\stackrel{}{r}$$
(35)
and the magnetic transition dipole
$$\stackrel{}{\mu }_{mn}=\frac{1}{2c}\stackrel{}{r}\times \stackrel{}{j}_{mn}(\stackrel{}{r})๐\stackrel{}{r}.$$
(36)
Using the equality
$$\stackrel{}{}=m_0[\stackrel{}{r},\frac{^2}{2m_0}+e\phi _i(\stackrel{}{r})],$$
one can connect the electric transition current (34) and transition dipole (35) as
$$\stackrel{}{j}_{mn}=i\omega _{mn}\stackrel{}{d}_{mn},\omega _{mn}E_mE_n.$$
(37)
The field operators can be expanded over the basis of the wave functions as
$$\psi (\stackrel{}{r},t)=\underset{n}{}\underset{i=1}{\overset{N}{}}c_{ni}(t)\psi _n(\stackrel{}{r}\stackrel{}{r}_i).$$
From the commutation relations for the field operators one has
$$[c_{mi}(t),c_{ni}^{}(t)]_{}=\delta _{mn}\delta _{ij},[c_{mi}(t),c_{nj}(t)]_{}=0.$$
The fact that each radiator is certainly in one of the states labelled by the index $`n`$ is expressed by the unipolarity condition
$$\underset{n}{}c_{ni}^{}(t)c_{ni}(t)=1.$$
(38)
The wave functions $`\psi _n(\stackrel{}{r}\stackrel{}{r}_i)`$, in agreement with inequalities (26), are localized in a small region of the size of a radiator. Such functions are called the localized orbitals. The localization condition can be represented by the equality
$$\psi _m^{}(\stackrel{}{r}\stackrel{}{r}_i)f(\stackrel{}{r})\psi _n(\stackrel{}{r}\stackrel{}{r}_j)๐\stackrel{}{r}=0(ij),$$
in which $`f(\stackrel{}{r})`$ is a finite function.
Using the notations and conditions introduced above, we transform the radiator Hamiltonian (30) to the form
$$\widehat{H}_r(t)=\underset{n}{}\underset{i=1}{\overset{N}{}}E_nc_{ni}^{}(t)c_{ni}(t).$$
(39)
The radiatorโfield Hamiltonian (32) becomes
$$\widehat{H}_{rf}(t)=\underset{mn}{}\underset{i=1}{\overset{N}{}}c_{mi}^{}(t)c_{ni}(t)\left[\stackrel{}{d}_{mn}\stackrel{}{E}_{0i}(t)+\frac{1}{c}\stackrel{}{j}_{mn}\stackrel{}{A}_i(t)+\stackrel{}{\mu }_{mn}\stackrel{}{B}_i(t)\right],$$
(40)
where
$$\stackrel{}{B}_i(t)=\stackrel{}{H}_{0i}(t)+\stackrel{}{H}_i(t)$$
(41)
is the total magnetic field.
From definitions (34) to (36), we have
$$\stackrel{}{d}_{mn}^{}=\stackrel{}{d}_{nm},\stackrel{}{j}_{mn}^{}=\stackrel{}{j}_{nm},\stackrel{}{\mu ^{}}_{mn}=\stackrel{}{\mu }_{nm}.$$
Because the wave functions are usually either symmetric or antisymmetric with respect to the spatial inversion, so that
$$|\psi _n(\stackrel{}{r})|=|\psi _n(\stackrel{}{r})|,$$
(42)
then we see that $`\stackrel{}{d}_{nn}=\stackrel{}{j}_{nn}=0`$ but, in general, $`\stackrel{}{\mu }_{nn}0`$.
The next approximation that is usually involved is related to the situation when only a couple of radiator levels takes part in the considered process. This happens when the transition frequency
$$\omega _0\omega _{21}=E_2E_1>0$$
(43)
for these two levels is selected by means of an external alternating field whose frequency is close to the transition frequency (43). In this way, considering only two levels is equivalent to the quasiresonance approximation. Then, it is convenient to introduce the transition operators
$$\sigma _i^{}(t)=c_{1i}^{}(t)c_{2i}(t),\sigma _i^+(t)=c_{2i}^{}(t)c_{1i}(t)$$
and the populationโdifference operator
$$\sigma _i^z(t)=c_{2i}^{}(t)c_{2i}(t)c_{1i}^{}(t)c_{1i}(t),$$
so that
$$2c_{1i}^{}(t)c_{1i}(t)=1\sigma _i^z(t),2c_{2i}^{}(t)c_{2i}(t)=1+\sigma _i^z(t).$$
The commutation relations for the introduced operators are
$$[\sigma _i^{},\sigma _j^+]=\delta _{ij}\sigma _i^z,[\sigma _i^{},\sigma _j^{}]=[\sigma _i^+,\sigma _j^+]=0,[\sigma _i^{},\sigma _j^z]=2\delta _{ij}\sigma _i^{},$$
$$[\sigma _i^+,\sigma _j^z]=2\delta _{ij}\sigma _i^+,[\sigma _i^{},\stackrel{}{A}_j]=[\sigma _i^{},\stackrel{}{E}_j]=[\sigma _i^z,\stackrel{}{A}_j]=[\sigma _i^z,\stackrel{}{E}_j]=0,$$
where all operators are taken at coinciding times.
With the notation
$$\stackrel{}{d}_{21}\stackrel{}{d},\stackrel{}{\mu }_{21}\stackrel{}{\mu },$$
(44)
we have $`\stackrel{}{d}_{12}=\stackrel{}{d}^{},\stackrel{}{\mu }_{12}=\stackrel{}{\mu ^{}}`$, and consequently
$$\stackrel{}{j}_{12}=i\omega _0\stackrel{}{d}^{},\stackrel{}{j}_{21}=i\omega _0\stackrel{}{d}.$$
(45)
Since only the difference between level energies is measurable, one can set $`E_1=0`$. Then the radiator Hamiltonian (39) reduces to
$$\widehat{H}_r(t)=\frac{1}{2}\underset{i=1}{\overset{N}{}}\omega _0\left[1+\sigma _i^z(t)\right].$$
(46)
Everywhere in what follows we assume that electromagnetic fields acting on a radiator do not change the classification of its energy levels. In the other case it would be impossible to talk about quasiresonance. This implies that the interaction energies of a radiator with fields are assumed to be much smaller than $`\omega _0`$. Because of the latter, the term
$$\frac{1}{2}\underset{i=1}{\overset{N}{}}\left[(\stackrel{}{\mu }_{11}+\stackrel{}{\mu }_{22})+(\stackrel{}{\mu }_{22}\stackrel{}{\mu }_{11})\sigma _i^z(t)\right]\stackrel{}{B}_i(t),$$
entering the radiatorโfield Hamiltonian (40), can be neglected as compared to Eq. (46). As a result, we obtain
$$\widehat{H}_{rf}(t)=\underset{i=1}{\overset{N}{}}\left[\frac{1}{c}\stackrel{}{j}_i(t)\stackrel{}{A}_i(t)+\stackrel{}{d}_i(t)\stackrel{}{E}_{0i}(t)+\stackrel{}{\mu }_i(t)\stackrel{}{B}_i(t)\right],$$
(47)
where the notation
$$\stackrel{}{j}_i(t)=i\omega _0\left[\stackrel{}{d}\sigma _i^+(t)\stackrel{}{d}^{}\sigma _i^{}(t)\right],\stackrel{}{d}_i(t)=\stackrel{}{d}\sigma _i^+(t)+\stackrel{}{d}^{}\sigma _i^{}(t),$$
$$\stackrel{}{\mu }_i(t)=\stackrel{}{\mu }\sigma _i^+(t)+\stackrel{}{\mu ^{}}\sigma _i^{}(t)$$
(48)
is used. The Hamiltonian of the matterโfield interaction can be written analogously to the first term in Eq. (47) as
$$\widehat{H}_{mf}(t)=\frac{1}{c}\underset{j=1}{\overset{N_0}{}}\stackrel{}{J}_{mj}(t)\stackrel{}{A}_j(t),$$
(49)
where $`N_0`$ is the number of particles forming the matter and $`\stackrel{}{J}_{mj}`$ is a local matter current having the structure of the operator $`\stackrel{}{J}_{mj}=(e/m)\stackrel{}{p}_j`$, with $`\stackrel{}{p}_j`$ being the momentum of a $`j`$โparticle.
The transition between the quantum states $`\psi _1`$ and $`\psi _2`$ can be either accompanied by the change of parity or not. Then from definitions (35) and (36) it follows that one has one of two possibilities:
$$\stackrel{}{d}0,\stackrel{}{\mu }=0(changedparity);$$
$$\stackrel{}{d}=0,\stackrel{}{\mu }0(conservedparity).$$
(50)
Thus, we actually have to deal with only one of the dipole transitions, either with electric or with magnetic. Here we consider them in parallel in order to compare these two cases.
## 4 Stochastic MeanโField Approximation
Now it is necessary to write down the evolution equations for the operators entering the total Hamiltonian (29) whose terms are given by Eqs (46), (31), (47), and (49). The Heinserberg equations yield
$$\frac{1}{c}\frac{}{t}\stackrel{}{E}(\stackrel{}{r},t)=\stackrel{}{}\times \stackrel{}{H}(\stackrel{}{r},t)\frac{4\pi }{c}\stackrel{}{J}(\stackrel{}{r},t),\frac{1}{c}\frac{}{t}\stackrel{}{A}(\stackrel{}{r},t)=\stackrel{}{E}(\stackrel{}{r},t),$$
(51)
which are, actually, the operator Maxwell equations, where the operator of current is
$$\stackrel{}{J}(\stackrel{}{r},t)=\underset{i=1}{\overset{N}{}}\left[\stackrel{}{j}_i(t)c\stackrel{}{\mu }_i(t)\times \stackrel{}{}\right]\delta (\stackrel{}{r}\stackrel{}{r}_i)+\underset{j=1}{\overset{N_0}{}}\stackrel{}{J}_{mj}(t)\delta (\stackrel{}{r}\stackrel{}{r}_j).$$
(52)
For the transition operators we have
$$\frac{d\sigma _i^{}}{dt}=i\omega _0\sigma _i^{}+\left(k_0\stackrel{}{d}\stackrel{}{A}_ii\stackrel{}{d}\stackrel{}{E}_{0i}i\stackrel{}{\mu }\stackrel{}{B}_i\right)\sigma _i^z$$
(53)
for the lowering operator, where $`k_0\omega _0/c`$, and the Hermitian conjugate equation for the rising operator $`\sigma _i^+`$. For the populationโdifference operator we get
$$\frac{d\sigma _i^z}{dt}=2k_0\left(\stackrel{}{d}\sigma _i^++\stackrel{}{d}^{}\sigma _i^{}\right)\stackrel{}{A}_i+2i\left(\stackrel{}{d}\sigma _i^+\stackrel{}{d}^{}\sigma _i^{}\right)\stackrel{}{E}_{0i}+2i\left(\stackrel{}{\mu }\sigma _i^+\stackrel{}{\mu ^{}}\sigma _i^{}\right)\stackrel{}{B}_i.$$
(54)
From Eqs. (51), using the Coulomb calibration, we find the wave equation
$$\left(\stackrel{}{}^2\frac{1}{c^2}\frac{^2}{t^2}\right)\stackrel{}{A}(\stackrel{}{r},t)=\frac{4\pi }{c}\stackrel{}{J}(\stackrel{}{r},t).$$
(55)
The solution of the latter has the form
$$\stackrel{}{A}(\stackrel{}{r},t)=\stackrel{}{A}_{vac}(\stackrel{}{r},t)+\frac{1}{c}\stackrel{}{J}(\stackrel{}{r}^{},t\frac{|\stackrel{}{r}\stackrel{}{r}^{}|}{c})\frac{d\stackrel{}{r}^{}}{|\stackrel{}{r}\stackrel{}{r}^{}|},$$
(56)
in which $`\stackrel{}{A}_{vac}`$ is the vacuum vector potential being a solution of the uniform wave equation. With the operator of current (52), the vector potential (56) can be written as the sum
$$\stackrel{}{A}=\stackrel{}{A}_{vac}+\stackrel{}{A}_{rad}+\stackrel{}{A}_{mat}$$
(57)
of the vacuum potential $`\stackrel{}{A}_{vac}`$, the radiator potential
$$\stackrel{}{A}_{rad}(\stackrel{}{r}_i,t)=\underset{j}{}\frac{1}{cr_{ij}}\stackrel{}{j}_j\left(t\frac{r_{ij}}{c}\right)+\underset{j}{}\frac{\stackrel{}{r}_{ij}}{r_{ij}^3}\times \left(r_{ij}\frac{}{r_{ij}}1\right)\stackrel{}{\mu }_j\left(t\frac{r_{ij}}{c}\right),$$
(58)
and of the matter potential
$$\stackrel{}{A}_{mat}(\stackrel{}{r}_i,t)=\underset{j}{}\frac{1}{cr_{ij}}\stackrel{}{J}_{mj}\left(t\frac{r_{ij}}{c}\right),$$
(59)
where $`\stackrel{}{r}_{ij}\stackrel{}{r}_i\stackrel{}{r}_j,r_{ij}|\stackrel{}{r}_{ij}|`$, and the summation $`_j`$ does not include the term with $`j=i`$.
Our aim is to derive the evolution equations for the variables
$$u_i(t)<\sigma _i^{}(t)>,s_i(t)<\sigma _i^z(t)>,$$
(60)
in which the angle brackets mean the statistical averaging over the radiator degrees of freedom. For the double correlators, we shall employ the meanโfield type decoupling
$$<\sigma _i^\alpha \sigma _j^\beta >=<\sigma _i^\alpha ><\sigma _j^\beta >(ij).$$
(61)
The quantum effects due to selfโaction can be taken into account by including into the evolution equations the attenuation terms defined by
$$\gamma \frac{4}{3}k_0^3\left(d_0^2+\mu _0^2\right),$$
(62)
where $`d_0|\stackrel{}{d}|`$ and $`\mu _0|\stackrel{}{\mu }|`$. More generally, one includes the phenomenological longitudinal and transverse attenuation parameters $`\gamma _1`$ and $`\gamma _2`$.
To take into account the retardation, we may remember that the action of electromagnetic fields is characterized by the energies that are much smaller than $`\omega _0`$. That is, in the zero order one has $`\sigma _i^{}\mathrm{exp}(i\omega _0t)`$, as follows from Eq. (53). This suggests to treat the retardation by means of the formula
$$<\sigma _j^{}\left(t\frac{r_{ij}}{c}\right)>=u_j(t)\mathrm{exp}(ik_0r_{ij}),$$
(63)
which can be called the quasirelativistic approximation since in the relativistic limit $`c\mathrm{}`$, Eq. (63) becomes an identity.
Comparing the terms of the vector potential (58), induced by either electrodipole or magnetodipole transitions, we notice their essential difference. Really, averaging over angles gives
$$\underset{j}{}f(r_{ij})\stackrel{}{r}_{ij}=0,$$
(64)
unless there is a special arrangement of radiators in space. Hence, the vector potential induced by magnetodipole transitions, in usual conditions, is negligibly small. Then for the averaged potential (58), we have
$$<\stackrel{}{A}_{rad}(\stackrel{}{r}_i,t)>=ik_0^2\underset{j}{}\left(\stackrel{}{d}\phi _{ij}^{}u_j^{}\stackrel{}{d}^{}\phi _{ij}u_j\right),$$
(65)
where
$$\phi _{ij}\frac{\mathrm{exp}(ik_0r_{ij})}{k_0r_{ij}}.$$
(66)
The influence of vacuum fluctuations and of matter is characterized by the term
$$\xi _i(t)k_0\stackrel{}{d}\left[\stackrel{}{A}_{vac}(\stackrel{}{r}_i,t)+\stackrel{}{A}_{mat}(\stackrel{}{r}_i,t)\right],$$
(67)
which we consider as a stochastic variable, whose properties are to be defined by additional conditions.
In this way, we come to the evolution equations for the transverse variable,
$$\frac{du_i}{dt}=(i\omega _0+\gamma _2)u_iis_i\left(\stackrel{}{d}\stackrel{}{E}_{0i}+\stackrel{}{\mu }\stackrel{}{H}_{0i}\right)+$$
$$+ik_0^3s_i\stackrel{}{d}\underset{j}{}\left(\stackrel{}{d}\phi _{ij}^{}u_j^{}\stackrel{}{d}^{}\phi _{ij}u_j\right)+s_i\xi _i,$$
(68)
and for the longitudinal variable,
$$\frac{ds_i}{dt}=2iu_i^{}\left(\stackrel{}{d}\stackrel{}{E}_{0i}+\stackrel{}{\mu }\stackrel{}{H}_{0i}\right)2iu_i\left(\stackrel{}{d}^{}\stackrel{}{E}_{0i}+\stackrel{}{\mu ^{}}\stackrel{}{H}_{0i}\right)$$
$$2ik_0^3(\stackrel{}{d}u_i^{}+\stackrel{}{d}^{}u_i)\underset{j}{}\left(\stackrel{}{d}\phi _{ij}^{}u_j^{}\stackrel{}{d}^{}\phi _{ij}u_j\right)\gamma _1(s_i\zeta )2(u_i^{}\xi _i+u_i\xi _i^{}),$$
(69)
where $`\zeta [1,1]`$ is a pumping parameter. An equation for $`u_i^{}`$ can be obtained by the complex conjugation of Eq. (68). Another useful equation is
$$\frac{d|u_i|^2}{dt}=2\gamma _2|u_i|^2+s_i(u_i^{}\xi _i+u_i\xi _i^{})is_iu_i^{}\left(\stackrel{}{d}\stackrel{}{E}_{0i}+\stackrel{}{\mu }\stackrel{}{H}_{0i}\right)+$$
$$+is_iu_i\left(\stackrel{}{d}^{}\stackrel{}{E}_{0i}+\stackrel{}{\mu ^{}}\stackrel{}{H}_{0i}\right)+ik_0^3s_i\left(u_i^{}\stackrel{}{d}+u_i\stackrel{}{d}^{}\right)\underset{j}{}\left(\stackrel{}{d}\phi _{ij}^{}u_j^{}\stackrel{}{d}^{}\phi _{ij}u_j\right).$$
(70)
Equations (68) and (70) are basic for describing nonequilibrium collective phenomena in radiating systems. The set of assumptions employed for deriving these equations can be briefly named the stochastic meanโfield approximation since the meanโfield type decoupling (61) was used for the radiator correlators, but quantum effects are taken into account through the stochastic variable (67).
## 5 Dynamical Characteristics of Coherence
One of the most important results of the cooperative behaviour of radiators is the appearance of coherent radiation. The level of coherence of electromagnetic fields can be described by the corresponding correlation functions . Here we introduce another characteristic of coherence, which is convenient for considering the radiation from ensembles of radiators .
The energy density of the radiated electromagnetic field is
$$W\frac{1}{8\pi }\left(\stackrel{}{E}^2+\stackrel{}{H}^2\right),$$
(71)
where $`\stackrel{}{E}=\stackrel{}{E}(\stackrel{}{r},t)`$ and $`\stackrel{}{H}=\stackrel{}{H}(\stackrel{}{r},t)`$. Differentiating Eq. (71) with respect to time, using the Maxwell equations (51), and defining the intensity of scattering
$$\frac{W_s}{t}\frac{1}{2}\left(\stackrel{}{J}\stackrel{}{E}+\stackrel{}{E}\stackrel{}{J}\right)$$
(72)
and the Poynting vector
$$\stackrel{}{S}\frac{c}{8\pi }\left(\stackrel{}{E}\times \stackrel{}{H}\stackrel{}{H}\times \stackrel{}{E}\right),$$
(73)
we obtain the continuity equation
$$\frac{}{t}\left(W+W_s\right)+\mathrm{div}\stackrel{}{S}=0.$$
(74)
The intensity of radiation into the unit solid angle is
$$I(\stackrel{}{n},t)<:\stackrel{}{n}\stackrel{}{S}(\stackrel{}{r},t):>r^2,$$
(75)
where $`\stackrel{}{n}\stackrel{}{r}/r,r|\stackrel{}{r}|`$, and the colons imply the normal ordering of operators. To accomplish the latter, one separates the Hermitian operators into their conjugate parts, which, for instance, for the vector potential (58) reads as
$$\stackrel{}{A}_{rad}(\stackrel{}{r},t)=\stackrel{}{A}^+(\stackrel{}{r},t)+\stackrel{}{A}^{}(\stackrel{}{r},t),$$
(76)
where
$$\stackrel{}{A}^+(\stackrel{}{r},t)=\underset{j}{}\left[\frac{ik_0\stackrel{}{d}}{|\stackrel{}{r}\stackrel{}{r}_j|}+\frac{1+ik_0|\stackrel{}{r}\stackrel{}{r}_j|}{|\stackrel{}{r}\stackrel{}{r}_j|^3}\stackrel{}{\mu }\times \left(\stackrel{}{r}\stackrel{}{r}_j\right)\right]\sigma _j^+\left(t\frac{1}{c}|\stackrel{}{r}\stackrel{}{r}_j|\right).$$
Respectively, the electromagnetic positive and negative fields related to Eq. (76) are
$$\stackrel{}{E}_{rad}\frac{1}{c}\frac{\stackrel{}{A}_{rad}}{t}=\stackrel{}{E}^++\stackrel{}{E}^{},\stackrel{}{H}_{rad}\stackrel{}{}\times \stackrel{}{A}_{rad}=\stackrel{}{H}^++\stackrel{}{H}^{}.$$
In the time and space derivatives, we may employ, for differentiating $`\sigma _j^\pm `$, the relations
$$\left(\frac{1}{c}\frac{}{t}+\frac{}{r_{ij}}\right)\sigma _j^\pm \left(t\frac{r_{ij}}{c}\right)=0,\left(\frac{}{r_{ij}}\pm ik_0\right)\sigma _j^\pm \left(t\frac{r_{ij}}{c}\right)=0.$$
In the wave zone, where $`r|\stackrel{}{r}_i|`$ and $`|\stackrel{}{r}\stackrel{}{r}_j|r\stackrel{}{n}\stackrel{}{r}_j,(r|\stackrel{}{r}_j|)`$, we have
$$\stackrel{}{A}^+(\stackrel{}{r},t)i\frac{k_0}{r}\left(\stackrel{}{d}+\stackrel{}{\mu }\times \stackrel{}{n}\right)\underset{j}{}\sigma _j^+\left(t\frac{r\stackrel{}{n}\stackrel{}{r}_j}{c}\right),$$
(77)
from where
$$\stackrel{}{E}^+=ik_0\stackrel{}{A}^+,\stackrel{}{H}^+=\stackrel{}{n}\times \stackrel{}{E}^+.$$
(78)
Then in the part of the Poynting vector (73), describing the radiation from the ensemble of radiators, one has
$$\stackrel{}{S}_{rad}=\frac{c}{4\pi }\stackrel{}{E}_{rad}\times \stackrel{}{H}_{rad},\stackrel{}{H}_{rad}=\stackrel{}{n}\times \stackrel{}{E}_{rad}.$$
For the corresponding part of the radiation intensity (75), we get
$$I_{rad}(\stackrel{}{n},t)=\frac{cr^2}{4\pi }<:\stackrel{}{E}_{rad}^2(\stackrel{}{n}\stackrel{}{E}_{rad})^2:>.$$
(79)
Averaging the latter over stochastic variables and over fast oscillations yields
$$\overline{I}(\stackrel{}{n},t)\frac{\omega _0}{2\pi }_0^{2\pi /\omega _0}I_{rad}(\stackrel{}{n},t)dt,$$
(80)
the slow variables in the process of integration being kept fixed. For the radiation intensity (79), this results in
$$\overline{I}(\stackrel{}{n},t)=\omega _0\gamma \underset{ij}{\overset{N}{}}f_{ij}(\stackrel{}{n})\overline{<\sigma _i^+(t)\sigma _j^{}(t)>},$$
(81)
where
$$f_{ij}(\stackrel{}{n})\frac{3}{8\pi }\left|\stackrel{}{n}\times \stackrel{}{e}\right|^2\mathrm{exp}\left(ik_0\stackrel{}{n}\stackrel{}{r}_{ij}\right)$$
(82)
and $`\stackrel{}{e}=\stackrel{}{d}/d_0`$ or $`\stackrel{}{\mu }/\mu _0`$ depending on the type of radiation.
In the radiation intensity (81), we may separate the terms with the coinciding and with different indices, so that $`_{ij}=_{i=j}+_{ij}`$. This makes it possible to separate the radiation intensity into the incoherent and coherent parts,
$$\overline{I}(\stackrel{}{n},t)=I_{inc}(\stackrel{}{n},t)+I_{coh}(\stackrel{}{n},t),$$
(83)
so that the incoherent radiation intensity is
$$I_{inc}(\stackrel{}{n},t)=\frac{1}{2}\omega _0\gamma \underset{i=1}{\overset{N}{}}f_{ii}(\stackrel{}{n})[1+s_i(t)]$$
(84)
while the coherent radiation intensity is
$$I_{coh}(\stackrel{}{n},t)=\omega _0\gamma \underset{ij}{\overset{N}{}}f_{ij}(\stackrel{}{n})\overline{u_i^{}(t)u_j(t)}.$$
(85)
Here the equality $`2\sigma _i^+\sigma _i^{}=1+\sigma _i^z`$ was used. The total radiation intensity is given by the integral
$$I(t)\overline{I}(\stackrel{}{n},t)๐\mathrm{\Omega }(\stackrel{}{n})=I_{inc}(t)+I_{coh}(t)$$
(86)
over solid angles. Here the incoherent part is
$$I_{inc}(t)=\frac{1}{2}\omega _0\gamma \underset{i=1}{\overset{N}{}}[1+s_i(t)],$$
(87)
and the coherent part is
$$I_{coh}(t)=\omega _0\gamma \underset{ij}{\overset{N}{}}f_{ij}\overline{u_i^{}(t)u_j(t)},$$
(88)
where
$$f_{ij}f_{ij}(\stackrel{}{n})๐\mathrm{\Omega }(\stackrel{}{n}),f_{ii}=1.$$
(89)
Finally, the level of coherence can be defined by means of the coherence coefficients
$$C_{coh}(\stackrel{}{n},t)\frac{I_{coh}(\stackrel{}{n},t)}{I_{inc}(\stackrel{}{n},t)},C_{coh}(t)\frac{I_{coh}(t)}{I_{inc}(t)}.$$
(90)
The radiation is mainly incoherent when $`C_{coh}1`$ and it is almost purely coherent if $`C_{coh}1`$.
## 6 Collective Liberation of Light
A system of initially inverted atoms can, due to photon exchange, become strongly correlated, as a result emitting a coherent pulse. This effect of selfโorganization, accompanied by a coherent burst, is called the Dicke superradiance . This phenomenon is well studied for atoms in vacuum , including different particular cases, such as superradiance in twoโcomponent systems \[35โ37\], superradiance from ensembles of threeโlevel molecules , twoโphoton superradiance , and so on (see citations in Refs. ). When radiating atoms or molecules are placed in a solid, they interact with phonons , which can lead to such interesting phenomena as the laser cooling of solids .
When an atom is placed in a periodic dielectric structure, in which, due to periodicity, a photonic band gap develops, then spontaneous emission with a frequency inside the band gap can be rigorously forbidden . This kind of matter, where photon band gap appears because of the structure periodicity in real space, has been called photonic bandโgap materials. The photon band gap also appears in natural dense media due to photon interactions with optical collective excitations, such as phonons, magnons, or excitons . One calls this type of the gap the polariton band gap since photons coupled with collective excitations of a medium are termed polaritons.
If a single resonance atom is placed in a medium with a photon band gap, and the atomic transition frequency lies inside this gap, then the spontaneous emission is suppressed, which is named the localization of light . This effect is caused by the formation of a photonโatom bound state \[50โ52\]. When a collection of identical resonance atoms is doped into a medium with a photon band gap, so that the atomic transition frequency is inside this gap, then the atoms, in principle, can radiate because of the formation of a photonic impurity band within the photon band gap \[50,53โ55\]. A model case of a concentrated sample, whose linear size $`L`$ is much smaller than the radiation wavelength $`\lambda `$, has been considered for studying superradiance near a photonic band gap , when the transition frequency almost coincides with the frequency of the upper band edge. Here, following Ref. , we study the realistic case of a sample with $`\lambda L`$.
Assume that the localization of light occurs for a single atom with an electric dipole transition, so that its population difference is always $`s_0=s(0)`$. Considering an ensemble of resonance atoms, we resort to Eqs. (68), (69), and (70). For simplicity, we write $`u_i=u`$ and $`s_i=s`$. Introduce the effective coupling parameters
$$g\frac{3\gamma }{4\gamma _2}\underset{j}{}\frac{\mathrm{sin}(k_0r_{ij})}{k_0r_{ij}},g^{}\frac{3\gamma }{4\gamma _2}\underset{j}{}\frac{\mathrm{cos}(k_0r_{ij})}{k_0r_{ij}},$$
(91)
where $`\gamma 4k_0^3d_0^2/3`$. In the absence of resonator imposing a selected mode,
$$gg^{}\frac{3\gamma }{4\gamma _2}\rho \lambda ^3,$$
(92)
where $`\rho `$ is the density of resonance atoms. It is convenient to introduce the effective frequency and effective attenuation defined, respectively, as
$$\mathrm{\Omega }\omega _0+\gamma _2g^{}s,\mathrm{\Gamma }\gamma _2(1gs).$$
(93)
These expressions include the influence of local fields through the coupling parameters (91). Since the latter take into account the existence of an ensemble of atoms, we may call $`\mathrm{\Omega }`$ and $`\mathrm{\Gamma }`$ the collective frequency and collective width, respectively.
With these notations, Eq. (68) reduces to
$$\frac{du}{dt}=\left(i\mathrm{\Omega }+\mathrm{\Gamma }\right)u+s\xi +\gamma _2\stackrel{}{e_d}^2(g+ig^{})su^{},$$
(94)
where $`\xi =\xi _i`$ and $`\stackrel{}{e}_d\stackrel{}{d}/d_0`$. Equation (69) becomes
$$\frac{ds}{dt}=4\gamma _2g|u|^2\gamma _1(ss_0)2(u^{}\xi +u\xi ^{})$$
$$2\gamma _2\left[(g+ig^{})\left(u^{}\stackrel{}{e}_d\right)^2+(gig^{})\left(u\stackrel{}{e_d}^{}\right)^2\right],$$
(95)
where $`\zeta =s_0`$ takes into account that for a single atom the localization of light occurs. And for Eq. (70), we have
$$\frac{d|u|^2}{dt}=2\mathrm{\Gamma }|u|^2+s\left(u^{}\xi +\xi ^{}u\right)+\gamma _2s\left[(g+ig^{})\left(u^{}\stackrel{}{e}_d\right)^2+(gig^{})\left(u\stackrel{}{e_d}^{}\right)^2\right].$$
(96)
Let us accept the natural inequalities
$$\frac{\gamma _1}{\mathrm{\Omega }}1,\frac{\gamma _2}{\mathrm{\Omega }}1,\left|\frac{\mathrm{\Gamma }}{\mathrm{\Omega }}\right|1.$$
(97)
And, as always, we keep in mind that the interaction term (67) is small as compared to the frequency $`\mathrm{\Omega }`$, or that $`\xi =0`$, which tells that this term is small on average. Then, according to Sec. 2, we may classify the solution $`u`$ as fast while $`s`$ and$`|u|^2`$ as slow. Solving Eq. (94), with $`s`$ being a quasiโinvariant, we get
$$u(t)=\left[u_0+s_0^te^{(i\mathrm{\Omega }+\mathrm{\Gamma })t^{}}\xi (t^{})๐t^{}\right]e^{(i\mathrm{\Omega }+\mathrm{\Gamma })t}.$$
(98)
Introduce the notation
$$\alpha \underset{\tau \mathrm{}}{lim}\frac{\mathrm{Re}}{\tau \mathrm{\Gamma }s}_0^\tau \xi ^{}(t)u(t)dt,$$
(99)
where $`\mathrm{Re}`$ means the real part and which, if $`\xi =0`$, takes the form
$$\alpha =\underset{\tau \mathrm{}}{lim}\frac{\mathrm{Re}}{\tau \mathrm{\Gamma }}_0^\tau ๐t_0^te^{(i\mathrm{\Omega }+\mathrm{\Gamma })(tt^{})}\xi ^{}(t)\xi (t^{})dt^{}.$$
When $`\xi (t)`$ is a stochastic variable corresponding to a stationary random process, so that
$$\xi ^{}(t)\xi (t^{})=\xi ^{}(tt^{})\xi (0),$$
then the notation (99) becomes
$$\alpha =\underset{\tau \mathrm{}}{lim}\frac{\mathrm{Re}}{\tau \mathrm{\Gamma }}_0^\tau ๐t_0^te^{(i\mathrm{\Omega }+\mathrm{\Gamma })t^{}}\xi ^{}(t^{})\xi (0)dt^{}.$$
Defining a new function
$$w|u|^2\alpha s^2,$$
(100)
and averaging the rightโhand sides of Eqs. (95) and (96) over time and over stochastic variables we get
$$\frac{ds}{dt}=4g\gamma _2w\gamma _1^{}(s\zeta ^{}),\frac{d|u|^2}{dt}=2\mathrm{\Gamma }w,$$
where
$$\gamma _1^{}\gamma _1+4\gamma _2\alpha ,\zeta ^{}\frac{\gamma _1}{\gamma _1^{}}s_0.$$
In what follows, we assume that the quantity (99), describing the intensity of interaction between atoms and matter, is small,
$$|\alpha |1.$$
(101)
To understand the structure of the atomโmatter coupling $`\alpha `$, we may model the random variable $`\xi `$ by the interaction of an atom with an ensemble of oscillators as
$$\xi (t)=\underset{\omega }{}\gamma _\omega \left(b_\omega e^{i\omega t}+b_\omega ^{}e^{i\omega t}\right),$$
where $`b_\omega `$ and $`b_\omega ^{}`$ are Bose operators. Then the atomโmatter coupling is
$$\alpha =\underset{\omega }{}\gamma _\omega ^2\left[\frac{n_\omega }{(\omega \mathrm{\Omega })^2+\mathrm{\Gamma }^2}+\frac{1+n_\omega }{(\omega +\mathrm{\Omega })^2+\mathrm{\Gamma }^2}\right],$$
with $`n_\omega b_\omega ^{}b_\omega `$. If the coupling $`\alpha `$ is small, then $`\gamma _1^{}\gamma _1,\zeta ^{}s_0`$, and $`d|u|^2/dtdw/dt`$. Therefore, we obtain the equations
$$\frac{ds}{dt}=4g\gamma _2w\gamma _1(ss_0),\frac{dw}{dt}=2\gamma _2(1gs)w.$$
(102)
For transient times, when $`t\gamma _1^1`$, Eqs. (102) can be solved explicitly, giving
$$s=\frac{\gamma _0}{g\gamma _2}\mathrm{tanh}\left(\frac{tt_0}{\tau _0}\right)+\frac{1}{g},w=\frac{\gamma _0^2}{4g^2\gamma _2^2}\mathrm{sech}^2\left(\frac{tt_0}{\tau _0}\right),$$
(103)
where the integration constants $`\gamma _0=\tau _0^1`$ and $`t_0`$ are defined by the initial conditions $`u(0)=u_0`$ and $`s(0)=s_0`$. For the radiation width $`\gamma _0`$, we get the equation
$$\gamma _0^2=\mathrm{\Gamma }_0^2+4g^2\gamma _2^2\left(|u_0|^2\alpha _0s_0^2\right),$$
(104)
where
$$\mathrm{\Gamma }_0\gamma _2(1gs_0),\gamma _0\frac{1}{\tau _0},\alpha _0\alpha (0).$$
For the delay time, we find
$$t_0=\frac{\tau _0}{2}\mathrm{ln}\left|\frac{\gamma _0\mathrm{\Gamma }_0}{\gamma _0+\mathrm{\Gamma }_0}\right|.$$
(105)
Introducing the critical coupling
$$\alpha _c\frac{(1gs_0)^2}{4g^2s_0^2}+\frac{|u_0|^2}{s_0^2},$$
(106)
we may rewrite the radiation width as
$$\gamma _0=2g|s_0|\gamma _2\sqrt{\alpha _c\alpha _0}.$$
(107)
In the case of only one atom, we have to set $`g=0`$. Then Eqs. (102) give
$$s=s_0,w=(|u_0|^2\alpha _0s_0^2)e^{2\gamma _2t}(g=0),$$
which means that the light is localized. But for an ensemble of atoms the radiation becomes possible.
To find out what happens at large times, when $`t\mathrm{}`$, we need to analyse the stationary solutions of Eqs. (102). There are two pairs of such solutions:
$$s_1^{}=s_0,w_1^{}=0$$
(108)
and
$$s_2^{}=\frac{1}{g},w_2^{}=\frac{\gamma _1(gs_01)}{4g^2\gamma _2}.$$
(109)
The stability analysis shows that the fixed point (108) is stable for $`gs_0<1`$ and unstable for $`gs_0>1`$, when the point (109) becomes stable. When $`gs_0<1`$, the stationary point (108) is a stable node, while that (109) is a saddle point. In the interval $`1<gs_01+\gamma _1/8\gamma _2`$, the fixed point (108) is a saddle point, and that (109) is a stable node. For $`gs_0>1+\gamma _1/8\gamma _2`$, the stationary solutions (108) correspond again to a saddle point, while the fixed point (109) becomes a stable focus. In the latter case, the pulsing regime of radiation is realized, with the asymptotic period between pulses
$$T_p=\frac{4\pi }{|\gamma _1^2+8(1gs_0)\gamma _1\gamma _2|^{1/2}}.$$
(110)
However, at finite times the radiation pulses are not periodic, so that the characteristic time (110) is an approximate period only for $`t\mathrm{}`$.
In this way, when a single atom cannot radiate because of the localization of light, an ensemble of atoms can emit coherent radiation, provided that the interaction between atoms is sufficiently strong, so that $`gs_0>1`$. This is why such an effect can be called the collective liberation of light. However, this liberation is not complete but only partial since $`s_2^{}>0`$.
## 7 Amplification by Nonresonant Fields
An essential enchancement of radiation can occur due to correlations between radiators, which results in the emission of a coherent pulse. In order that these correlations could be sufficiently strong, it is usually required that the radiation wavelength would be much larger than the mean distance between radiators. If the latter is not the case, it is hardly probable that the selfโorganized coherence can develop. How would it be possible to amplify the radiation intensity for a system of radiators whose wavelenght is smaller than or comparable with the mean distance between them? This question is of high importance for shortโwave emission such as $`x`$โray and $`\gamma `$โray radiation. Coherent transient effects due to phase modulation of recoilless $`\gamma `$ radiation have been considered both theoretically and experimentally \[60โ63\]. A regenerated signal of gamma echo has been observed , which is similar to photon echo in optics . In the present section we explore the conditions when stationary enchancement of shortโwave radiation is feasible, being due to external nonresonant fields. Some preliminary results on the problem have been reported \[66โ68\], based on simplified models. Here the problem is considered more accurately, using the main Eqs. (68) to (70). The latter, in the case of shortโwave radiation, when the interaction of radiators can be neglected, take the form
$$\frac{du_i}{dt}=(i\omega _0+\gamma _2)u_iis_i\stackrel{}{d}\stackrel{}{E}_{0i},$$
(111)
$$\frac{ds_i}{dt}=2i(u_i^{}\stackrel{}{d}u_i\stackrel{}{d}^{})\stackrel{}{E}_{0i}\gamma _1(s_i\zeta ),$$
(112)
$$\frac{d|u_i|^2}{dt}=2\gamma _2|u_i|^2is_i(u_i^{}\stackrel{}{d}u_i\stackrel{}{d}^{})\stackrel{}{E}_{0i}.$$
(113)
The initial conditions are $`u_i(0)=u_0`$ and $`s_i(0)=s_0`$.
Assuming, as usual, the existence of small parameters
$$\frac{\gamma _1}{\omega _0}1,\frac{\gamma _2}{\omega _0}1,\frac{|\stackrel{}{d}\stackrel{}{E}_{0i}|}{\omega _0}1,$$
(114)
we see that $`u_i`$ has to be classified as a fast solution while $`s_i`$ and $`|u_i|^2`$, as slow ones. With $`s_i`$ being a quasiโinvariant, Eq. (111) gives
$$u_i(t)=e^{(i\omega _0+\gamma _2)t}\left[u_0is_i\stackrel{}{d}_0^t\stackrel{}{E}_{0i}(\tau )e^{(i\omega _0+\gamma _2)\tau }๐\tau \right].$$
Let the external field $`\stackrel{}{E}_{0i}=\stackrel{}{E}_{0i}(t)`$ consist of two parts,
$$\stackrel{}{E}_{0i}=\stackrel{}{E}_0+\stackrel{}{E}_1e^{i(\stackrel{}{k}\stackrel{}{r}_i\omega t)}+\stackrel{}{E}_1^{}e^{(\stackrel{}{k}\stackrel{}{r}_i\omega t)},$$
(115)
one being a stationary nonresonant field $`\stackrel{}{E}_0`$, and another part is a pair of plane waves, which are in quasiresonance with the transition frequency,
$$\frac{|\mathrm{\Delta }|}{\omega _0}1,\mathrm{\Delta }\omega \omega _0.$$
(116)
Then the solution of Eq. (111) writes
$$u_i(t)=\frac{s_i\stackrel{}{d}\stackrel{}{E}_0}{\omega _0i\gamma _2}+\frac{s_i\stackrel{}{d}\stackrel{}{E}_1}{\mathrm{\Delta }+i\gamma _2}e^{i(\stackrel{}{k}\stackrel{}{r}_i\omega t)}+$$
$$+\left(u_0+\frac{s_i\stackrel{}{d}\stackrel{}{E}_0}{\omega _0i\gamma _2}\frac{s_i\stackrel{}{d}\stackrel{}{E}_1}{\mathrm{\Delta }+i\gamma _2}e^{i\stackrel{}{k}\stackrel{}{r}_i}\right)e^{(i\omega _0+\gamma _2)t}.$$
(117)
Substituting this into the rightโhand side of Eq. (112) and averaging over time as
$$\underset{\tau \mathrm{}}{lim}\frac{1}{\tau }_0^\tau f(s,t)๐t,$$
we come to the equation
$$\frac{ds_i}{dt}=\gamma _1^{}(s_i\zeta ^{}),$$
(118)
with
$$\gamma _1^{}\gamma _1+4\gamma _2\left(\frac{|\stackrel{}{d}\stackrel{}{E}_0|^2}{\omega _0^2+\gamma _0^2}+\frac{|\stackrel{}{d}\stackrel{}{E}_1|^2}{\mathrm{\Delta }^2+\gamma _2^2}\right),\zeta ^{}\frac{\gamma _1}{\gamma _1^{}}\zeta .$$
The solution to Eq. (118) is
$$s_i(t)=s_0e^{\gamma _1^{}t}+\zeta ^{}\left(1e^{\gamma _1^{}t}\right).$$
(119)
Calculating the correlation function
$$\overline{u_i^{}(t)u_j(t)}=s^2(t)\left(\frac{|\stackrel{}{d}\stackrel{}{E}_0|^2}{\omega _0^2+\gamma _2^2}+\frac{|\stackrel{}{d}\stackrel{}{E}_1|^2}{\mathrm{\Delta }^2+\gamma _2^2}e^{i\stackrel{}{k}\stackrel{}{r}_{ij}}\right),$$
where, for simplicity, we set $`s_i=s`$, we find the incoherent and coherent radiation intensities (84) and (85), respectively, as
$$I_{inc}(\stackrel{}{n},t)=\frac{3N}{16\pi }\omega _0\gamma \left|\stackrel{}{n}\times \stackrel{}{e}_d\right|^2[1+s(t)],$$
$$I_{coh}(\stackrel{}{n},t)=\frac{3N^2}{8\pi }\omega _0\gamma |\stackrel{}{n}\times \stackrel{}{e}_d|^2s^2(t)\times $$
$$\times \left[F(k_0\stackrel{}{n})\frac{|\stackrel{}{d}\stackrel{}{E}_0|^2}{\omega _0^2+\gamma _2^2}+F(k_0\stackrel{}{n}\stackrel{}{k})\frac{|\stackrel{}{d}\stackrel{}{E}_1|^2}{\mathrm{\Delta }^2+\gamma _2^2}\right],$$
(120)
where $`\stackrel{}{n}\stackrel{}{r}/r`$ and the form factor is
$$F(\stackrel{}{k})\frac{1}{N^2}\underset{ij}{\overset{N}{}}e^{i\stackrel{}{k}\stackrel{}{r}_{ij}}=\left|\frac{1}{N}\underset{i=1}{\overset{N}{}}e^{i\stackrel{}{k}\stackrel{}{r}_i}\right|^2.$$
(121)
As is seen from expressions (120) and (121), the maxima of coherent radiation occur in the directions satisfying the condition
$$\left(k_0\stackrel{}{n}\stackrel{}{k}\right)\stackrel{}{r}_i=2\pi n_i(n_i=0,1,2,\mathrm{}).$$
(122)
This corresponds either to forward scattering, when all $`n_i=0`$, and the periodicity of matter is not required, or to the scattering in the Bragg directions, for which the strict space periodicity of radiators is needed. The enhancement of coherent radiation in the directions defined by condition (122) is called the Borrmann effect , which for the case of $`\gamma `$โrays is sometimes termed the KaganโAfanasiev effect .
The total radiation intensities (87) and (88) are
$$I_{inc}(t)=\frac{1}{2}N\omega _0\gamma [1+s(t)],$$
$$I_{coh}(t)=N^2\phi \omega _0\gamma s^2(t)\left(\frac{|\stackrel{}{d}\stackrel{}{E}_0|^2}{\omega _0^2+\gamma _2^2}+\frac{|\stackrel{}{d}\stackrel{}{E}_1|^2}{\mathrm{\Delta }^2+\gamma _2^2}\right),$$
(123)
where the shape factor is
$$\phi \frac{3}{8\pi }\left|\stackrel{}{n}\times \stackrel{}{e}_d\right|^2F(k_0\stackrel{}{n}\stackrel{}{k})๐\mathrm{\Omega }(\stackrel{}{n}).$$
(124)
The value of the latter strongly depends on the shape of the considered sample. Thus, for pencilโlike or diskโlike shapes , one has
$`\phi =\{\begin{array}{ccc}\frac{3\lambda }{8L},& \frac{\lambda }{2\pi L}1,& \frac{R}{L}1\\ & & \\ \frac{3}{8}\left(\frac{\lambda }{\pi R}\right)^2,& \frac{\lambda }{2\pi R}1,& \frac{L}{R}1,\end{array}`$
where $`R`$ and $`L`$ are the radius and length of a cylindrical sample, and $`\lambda 2\pi /k,k|\stackrel{}{k}|=\omega /c`$.
Consider the stationary limit $`t\mathrm{}`$, keeping in mind the situation typical of Mรถsbauer experiments, when the alternating field is weak,
$$\frac{|\stackrel{}{d}\stackrel{}{E}_1|^2}{\gamma _1\gamma _2}1,$$
(128)
and let us set, for simplicity,
$$\zeta =1$$
(129)
which means that there is no additional pumping except through the given field (115). Then Eq. (119) reduces to
$$\underset{t\mathrm{}}{lim}s_i(t)=1+\frac{4\gamma _2}{\gamma _1}\left(\frac{|\stackrel{}{d}\stackrel{}{E}_0|^2}{\omega _0^2}+\frac{|\stackrel{}{d}\stackrel{}{E}_1|^2}{\mathrm{\Delta }^2+\gamma _2^2}\right).$$
For the coherence coefficient, defined in Eq. (90), we get
$$\underset{t\mathrm{}}{lim}C_{coh}(t)=N\frac{\phi \gamma _1}{2\gamma _2}.$$
(130)
The role of the nonresonant field $`\stackrel{}{E}_0`$ can be characterized by the switching factor
$$S(E_0,t)\frac{I(t)}{lim_{E_00}I(t)}$$
(131)
and its stationary limit
$$S(E_0)\underset{t\mathrm{}}{lim}S(E_0,t).$$
(132)
For our case, we obtain
$$S(E_0)=1+\frac{\mathrm{\Delta }^2+\gamma _2^2}{\omega _0^2}\left|\frac{\stackrel{}{d}\stackrel{}{E}_0}{\stackrel{}{d}\stackrel{}{E}_1}\right|^2.$$
(133)
The switching factors (128) and (129) show how the radiation intensity is amplified when a nonresonant field $`\stackrel{}{E}_0`$ is switched on, as compared to the situation when $`\stackrel{}{E}_0=0`$. As is seen from expression (130), the amplification can be quite noticeable only if $`|\stackrel{}{d}\stackrel{}{E}_0||\stackrel{}{d}\stackrel{}{E}_1|`$, so that to compensate the smallness of the parameters $`|\mathrm{\Delta }|/\omega _0`$ and $`\gamma _2/\omega _0`$.
## 8 Mรถssbauer Magnetic Anomaly
Stationary fields, electric or magnetic, can be due not to external sources but can arise in a sample as a result of phase transitions . If an ensemble of radiators is incorporated into matter exhibiting a phase transition accompanied by the appearance of a constant field, the latter may influence some radiation characteristics. An interesting example of this kind is given by the gamma radiation of Mรถssbauer nuclei placed into magnetic materials. This example is especially intriguing because of longโstanding controversy related to its interpretation.
There exists a number of experiments demonstrating the soโcalled magnetic anomaly of the Mรถssbauer effect in materials undergoing magnetic phase transition. This anomaly consists in an essential increase, up to $`50\%`$, of the area under the Mรถssbauer spectrum below the temperature of magnetic transition, as compared to the spectrum area in paramagnetic state above the transition temperature. A detailed discussion of these experiments can be found in the book and review . The controversy related to this anomaly concerns the explanation of the cause of the latter.
The area of the Mรถssbauer spectrum, for Mรถssbauer nuclei in a solid sample, is given by the integral
$$A_{abs}=f_M_{\mathrm{}}^+\mathrm{}\sigma _{abs}(\omega )๐\omega ,$$
(134)
in which
$$f_M=\mathrm{exp}(k_0^2r_0^2)$$
(135)
is the Mรถssbauer factor, $`k_0=\omega _0/c,r_0`$ is the meanโsquare deviation of the nucleus from a lattice site,
$$\sigma _{abs}(\omega )=\frac{\sigma _0\mathrm{\Gamma }_{abs}^2}{(\omega \omega _0)^2+\mathrm{\Gamma }_{abs}^2}$$
(136)
is the absorption crossโsection, $`\mathrm{\Gamma }_{abs}`$ is the absorption halfโwidth,
$$\sigma _0=\frac{2\pi (1+2I_1)}{k_0^2(1+2I_0)(1+\alpha _e)}$$
(137)
is the crossโsection of resonant absorption, $`I_0`$ and $`I_1`$ are the nuclear spins of the groundโstate and excited levels, and $`\alpha _e`$ is the electron conversion coefficient. After integrating Eq. (131), we have the spectrum area
$$A_{abs}=\pi f_M\sigma _0\mathrm{\Gamma }_{abs}.$$
(138)
It is important to emphasize that the Mรถssbauer anomaly, we consider here, has been observed only in the soโcalled absorption geometry, when absorbing Mรถssbauer nuclei are placed inside magnetic matter which is irradiated by an external source. Contrary to this, in the experiments with the soโcalled source geometry, when a radioactive source is incorporated into the magnetic matter, but absorbing Mรถssbauer nuclei are outside this matter, no magnetic anomaly has been observed \[77โ79\]. Therefore it is clear that the considered Mรถssbauer anomaly is directly related to the action on Mรถssbauer nuclei of an effective magnetic field appearing below the critical point. But what is the origin of this anomaly?
Historically, the first suggestion was to ascribe the anomaly in the temperature behaviour of the spectrum area (135) to the influence of the appearing magnetic order on the Mรถssbauer factor (132). A number of citations having to do with this suggestion are listed in Refs. . This assumption implies that the meanโsquare deviation $`r_0`$ defining the Mรถssbauer factor (132) is essentially influenced by arising magnetic order. The course of reasoning is as follows. Mรถssbauer nuclei doped into a solid are characterized by the same meanโsquare deviation as the particles forming the solid sample. The latter can be described by the Hamiltonian
$$\widehat{H}_m=\underset{i}{}\frac{\stackrel{}{p_i}^2}{2m}+\frac{1}{2}\underset{ij}{}\mathrm{\Phi }(R_{ij})\underset{ij}{}I(R_{ij})\stackrel{}{S}_i\stackrel{}{S}_j,$$
(139)
in which $`\mathrm{\Phi }(R_{ij})`$ is a potential of direct pair interactions while $`I(R_{ij})`$ is that of exchange interactions, $`\stackrel{}{S}_i`$ is a spin operator, and $`R_{ij}|\stackrel{}{R}_{ij}|`$, with $`\stackrel{}{R}_{ij}=\stackrel{}{R}_i\stackrel{}{R}_j`$. The indices of summation in Eq. (136) run as $`i=1,2,\mathrm{},N_0`$, with $`N_0`$ being the number of lattice sites. Introduce the deviation from a lattice site,
$$\stackrel{}{u}_i\stackrel{}{R}_i\stackrel{}{a}_i,$$
(140)
defined so that
$$\stackrel{}{a}_i=<\stackrel{}{R}_i>,<\stackrel{}{u}_i>=0.$$
(141)
Taking into account that $`|\stackrel{}{u}_i|`$ is small as compared to the interparticle distance, one expands the interaction potential in powers of $`u_i^\alpha `$ up to the second order, which results in the Hamiltonian
$$\widehat{H}_m=U_0+\widehat{H}_p+\widehat{H}_s+\widehat{H}_{sp}+\widehat{H}^{},$$
(142)
whose terms are explained below: the constant part of the lattice energy
$$U_0=\frac{1}{2}\underset{ij}{}\mathrm{\Phi }(a_{ij});a_{ij}|\stackrel{}{a}_{ij}|,\stackrel{}{a}_{ij}\stackrel{}{a}_i\stackrel{}{a}_j,$$
(143)
the phonon term
$$\widehat{H}_p=\underset{i}{}\frac{\stackrel{}{p_i}^2}{2m}+\frac{1}{2}\underset{ij}{}\underset{\alpha \beta }{}\mathrm{\Phi }_{ij}^{\alpha \beta }u_i^\alpha u_j^\beta ,$$
(144)
in which $`\mathrm{\Phi }_{ij}^{\alpha \beta }^2\mathrm{\Phi }(a_{ij})/a_i^\alpha a_j^\beta `$, the spin Hamiltonian
$$\widehat{H}_s=\underset{ij}{}I(a_{ij})S_{ij};S_{ij}\stackrel{}{S}_i\stackrel{}{S}_j,$$
(145)
the term responsible for spinโphonon interactions,
$$\widehat{H}_{sp}=\underset{ij}{}\underset{\alpha \beta }{}I_{ij}^{\alpha \beta }u_i^\alpha u_j^\beta S_{ij},$$
(146)
where $`I_{ij}^{\alpha \beta }=^2I(a_{ij})/a_i^\alpha a_j^\beta `$, and the term
$$\widehat{H}^{}=\underset{i}{}\underset{\alpha }{}u_i^\alpha \left(1+\frac{1}{2}\underset{\beta }{}u_i^\beta \frac{}{a_i^\beta }\right)F_i^\alpha $$
(147)
related to the striction energy, where the striction force acting on the site $`i`$ is given by the components
$$F_i^\alpha \frac{}{a_i^\alpha }\underset{j}{}\left[\mathrm{\Phi }(a_{ij})2I(a_{ij})S_{ij}\right].$$
The correct definition of the lattice sites in Eq. (138) presupposes that they serve as equilibrium positions for particles. This implies that the striction energy is to be zero on average,
$$<\widehat{H}^{}>=0.$$
(148)
Then one invokes a kind of the semiclassical approximation
$$<u_i^\alpha S_{ij}>=<u_i^\alpha ><S_{ij}>=0,<u_i^\alpha u_j^\beta S_{ij}>=<u_i^\alpha u_j^\beta ><S_{ij}>,$$
decoupling the phonon and spin degrees of freedom, which suggests to present the operator term in the spinโphonon interaction (143) as
$$u_i^\alpha u_j^\beta S_{ij}=<u_i^\alpha u_j^\beta >S_{ij}+u_i^\alpha u_j^\beta <S_{ij}><u_i^\alpha u_j^\beta ><S_{ij}>.$$
(149)
Thus, the matter Hamiltonian (139) can be reduced to
$$\widehat{H}_m=\overline{U}_0+\widehat{\overline{H}}_p+\widehat{\overline{H}}_s,$$
(150)
with the renormalized terms
$$\overline{U}_0=U_0+\underset{ij}{}\underset{\alpha \beta }{}I_{ij}^{\alpha \beta }<u_i^\alpha u_j^\beta ><S_{ij}>,$$
$$\widehat{\overline{H}}_p=\underset{i}{}\frac{\stackrel{}{p_i}^2}{2m}+\frac{1}{2}\underset{ij}{}\underset{\alpha \beta }{}D_{ij}^{\alpha \beta }u_i^\alpha u_j^\beta ,\widehat{\overline{H}}_s=\underset{ij}{}J_{ij}S_{ij},$$
in which the striction energy, because of condition (145), is omitted and the renormalized interactions are
$$D_{ij}^{\alpha \beta }\mathrm{\Phi }_{ij}^{\alpha \beta }2I_{ij}^{\alpha \beta }<S_{ij}>,J_{ij}I(a_{ij})+\underset{\alpha \beta }{}I_{ij}^{\alpha \beta }<u_i^\alpha u_j^\beta >.$$
The renormalized dynamical matrix $`D_{ij}^{\alpha \beta }`$ defines the effective phonon spectrum $`\omega _{ks}`$ through the eigenvalue problem
$$\frac{1}{m}\underset{j}{}\underset{\beta }{}D_{ij}^{\alpha \beta }e^{\stackrel{}{k}\stackrel{}{a}_{ij}}e_{ks}^\beta =\omega _{ks}^2e_{ks}^\alpha ,$$
where $`\stackrel{}{e}_{ks}`$ is a polarization vector, the index $`s`$ labelling polarizations. The spectrum and polarization vectors are assumed to be even functions of the wave vector, so that $`\omega _{ks}=\omega _{ks}`$ and $`\stackrel{}{e}_{ks}=\stackrel{}{e}_{ks}`$. Polarization vectors form a complete orthonormal basis with the properties
$$\stackrel{}{e}_{ks}\stackrel{}{e}_{ks^{}}=\delta _{ss^{}},\underset{s}{}e_{ks}^\alpha e_{ks}^\beta =\delta _{\alpha \beta }.$$
Expanding the deviation and momentum as
$$\stackrel{}{u}_i=\underset{ks}{}\frac{\stackrel{}{e}_{ks}}{\sqrt{2mN_0\omega _{ks}}}\left(b_{ks}+b_{ks}^{}\right)e^{i\stackrel{}{k}\stackrel{}{a}_i},\stackrel{}{p}_i=i\underset{ks}{}\sqrt{\frac{m\omega _{ks}}{2N_0}}\stackrel{}{e}_{ks}\left(b_{ks}b_{ks}^{}\right)e^{i\stackrel{}{k}\stackrel{}{a}_i},$$
one transforms the renormalized phonon Hamiltonian to the standard form
$$\widehat{\overline{H}}_p=\underset{ks}{}\left(b_{ks}^{}b_{ks}+\frac{1}{2}\right).$$
After this, it is straightforward to calculate the correlators
$$<u_i^\alpha u_j^\beta >=\frac{\delta _{ij}}{2N_0}\underset{ks}{}\frac{e_{ks}^\alpha e_{ks}^\beta }{m\omega _{ks}}\mathrm{coth}\frac{\omega _{ks}}{2T},$$
in which $`T`$ is temperature. Thus, one gets the meanโsquare deviation from the equation
$$r_0^2\frac{1}{3}\underset{\alpha }{}<u_i^\alpha u_i^\alpha >=\frac{1}{6mN_0}\underset{ks}{}\frac{1}{\omega _{ks}}\mathrm{coth}\frac{\omega _{ks}}{2T}.$$
(151)
In this way, the influence of magnetic order on the meanโsquare deviation comes from its influence on the phonon spectrum.
However, the magnitude of the spinโphonon interaction, renormalizing the dynamical matrix, is rather small, as compared to the magnitude of direct interactions , so that $`|I_{ij}^{\alpha \beta }/\mathrm{\Phi }_{ij}^{\alpha \beta }|10^3`$. Hence, magnetic order cannot influence much phonon frequencies, as well as the sound velocity
$$c_s\underset{k0}{lim}\frac{\omega _{ks}}{k}=\underset{k0}{lim}\underset{j}{}\underset{\alpha \beta }{}D_{ij}^{\alpha \beta }\frac{(\stackrel{}{k}\stackrel{}{a}_{ij})^2}{2mk^2}e_{ks}^\alpha e_{ks}^\beta .$$
(152)
This conclusion is in agreement with all known experiments where phonon characteristics have been examined by means of neutron scattering, soundโvelocity measurements, elastic and thermal investigations. The onset of magnetic order can change the Mรถssbauer factor not more than by $`1\%`$, which cannot explain the observed Mรถssbauer anomaly of the spectrum area (135).
Another explanation was advanced by Babikova et al. , supposing that magnetic order can influence the electron conversion coefficient $`\alpha _e`$ in the crossโsection (134). A noticeable decrease of the conversion coefficient could lead to the increase of the crossโsection (134), and, consequently, to the increase of the spectrum area (135). The decrease of the conversion coefficient could be due to the suppression of the conversion channel in favour of the $`\gamma `$โradiation channel whose weight could be increased by the enhancement of the $`\gamma `$โradiation caused by the arising magnetic order .
To estimate the influence of an effective magnetic field, appearing in magnets, on the radiation intensity of Mรถssbauer nuclei, we have to consider the switching factor (130) that in our case, takes the form
$$S(H_0)=1+\frac{\gamma _2^2}{\omega _0^2}\left|\frac{\stackrel{}{\mu }\stackrel{}{H}_0}{\stackrel{}{\mu }\stackrel{}{H}_1}\right|^2.$$
For the characteristic Mรถssbauer nucleus $`{}_{}{}^{57}Fe`$, we have $`\omega _0=1.44\times 10^4`$eV and $`\gamma _2=\gamma _1=0.67\times 10^8`$eV, which can be transformed to the frequency units as $`\omega _010^{19}`$s<sup>-1</sup> and $`\gamma _2\gamma _110^7`$s<sup>-1</sup>. The corresponding wavelength is $`\lambda 10^8`$cm. Let us take for the effective magnetic field $`H_010^5`$G and for the alternating source field $`H_110^5`$G. The transition magnetic dipole $`\mu _00.1\mu _n`$, where $`\mu _n`$ is the nuclear magneton, hence $`\mu _010^{13}`$eV/G. This gives $`\mu _0H_010^7`$s<sup>-1</sup> and $`\mu _0H_110^3`$s<sup>-1</sup>. From here we obtain $`\gamma _2^2H_0^2/\omega _0^2H_1^210^4`$, which tells us that the switching factor $`S(H_0)`$ changes too little. Therefore, although the arising magnetic order does enhance the radiation of Mรถssbauer nuclei, this enhancement is not sufficient for causing such a drastic increase of the spectrum area.
The last quantity that could be blamed to be responsible for the Mรถssbauer magnetic anomaly is the absorption width $`\mathrm{\Gamma }_{abs}`$. The latter can be presented as the sum
$$\mathrm{\Gamma }_{abs}=\gamma _2+\gamma _2^{}$$
(153)
of the homogeneous line width $`\gamma _2`$ and of the inhomogeneous line width $`\gamma _2^{}`$. The inhomogeneous width can be due to the variation of local magnetic fields resulting in the random shift of the Mรถssbauer transition frequency . Returning to Section 3, we see that, really, an external magnetic field shifts the transition frequency as $`\omega _0+\left(\stackrel{}{\mu }_{22}\stackrel{}{\mu }_{11}\right)\stackrel{}{H}_0`$. Therefore, the inhomogeneous width can be of order $`\gamma _2^{}\left(\stackrel{}{\mu }_{22}\stackrel{}{\mu }_{11}\right)\stackrel{}{H}_0`$ or $`\gamma _2^{}\mu _0H_0`$. From here, $`\gamma _2^{}10^7`$s<sup>-1</sup>, that is, $`\gamma _2^{}\gamma _2`$. In this way, the anomalous increase of the Mรถssbauer spectrum area (135) below the magnetic transition temperature can be explained by the increase of the absorption width (150) caused by the increasing inhomogeneous width $`\gamma _2^{}\mu _0H_0`$.
## 9 Problem of Pattern Selection
Nonequilibrium cooperative phenomena are often described by nonlinear differential or integroโdifferential equations in partial derivatives. The solutions to such equations are in many cases nonuniform in space exhibiting the formation of different spatial structures. It happens that a given set of equations possesses several solutions corresponding to different spatial patterns . In such a case, the question arises which of these solutions, and respectively patterns, to prefer? The problem of pattern selection has no general solution . A possible way of selecting spatial structures, by minimizing the average energy, was delineated in subsection 2.5. Here we advance another method of pattern selection.
Assume that the considered differential equations in partial derivatives can be reduced to a $`d`$โdimensional system of ordinary equations; the dimensionality $`d`$ may equal infinity. Suppose also that admissible patterns are parametrized by a multiparameter $`\beta `$. Let the state of the dynamical system be defined by the set
$$y(t)=\{y_i(t)=y_i(\beta ,t)|i=1,2,\mathrm{},d\}$$
(154)
of solutions to the system of differential equations
$$\frac{d}{dt}y(t)=v(y,t).$$
(155)
For different parameters $`\beta `$ there are different sets (151) corresponding to different spatial structures. All admissible values of $`\beta `$ form a manifold $`=\{\beta \}`$. Each particular value of $`\beta `$ can be considered as a realization of the random variable from the manifold $``$. The classification of the states (151) can be done by defining a probability measure on $``$.
To introduce the probability distribution $`p(\beta ,t)`$ of patterns at time $`t`$, we resort to the ideas of statistical mechanics , where a probability $`p`$ can be connected with entropy $`S`$ by the relation $`pe^S`$. The entropy at time $`t`$ may be expressed as
$$S(t)\mathrm{ln}|\mathrm{\Delta }\mathrm{\Phi }(t)|$$
(156)
through the elementary phase volume
$$\mathrm{\Delta }\mathrm{\Phi }(t)\underset{i}{}\delta y_i(t).$$
(157)
Let us count the entropy from its initial value $`S(0)`$, thus, considering the entropy variation
$$\mathrm{\Delta }S(t)S(t)S(0).$$
(158)
Then the probability distribution $`pe^{\mathrm{\Delta }S}`$, normalized by the condition
$$p(\beta ,t)๐\beta =1$$
takes the form
$$p(\beta ,t)=\frac{e^{\mathrm{\Delta }S(\beta ,t)}}{Z(t)},$$
(159)
where the normalization factor is
$$Z(t)=e^{\mathrm{\Delta }S(\beta ,t)}๐\beta .$$
The entropy variation (155) writes
$$\mathrm{\Delta }S(t)=\mathrm{ln}\left|\frac{\mathrm{\Delta }\mathrm{\Phi }(t)}{\mathrm{\Delta }\mathrm{\Phi }(0)}\right|,$$
(160)
where the dependence on $`\beta `$, for brevity, is omitted. Define the multiplier matrix
$$M(t)=[M_{ij}(t)],M_{ij}(t)\frac{\delta y_i(t)}{\delta y_j(0)},$$
(161)
for which at the initial time one has
$$M_{ij}(0)\frac{\delta y_i(0)}{\delta y_j(0)}=\delta _{ij}.$$
(162)
The variation of the state (151) gives
$$\delta y(t)=M(t)\delta y(0),$$
(163)
which yields for the elementary phase volume (154)
$$\mathrm{\Delta }\mathrm{\Phi }(t)=\underset{i}{}\underset{j}{}M_{ij}(t)\delta y_j(0).$$
Hence, the entropy variation (157) is
$$\mathrm{\Delta }S(t)=\mathrm{ln}\left|\underset{i}{}\underset{j}{}M_{ij}(t)M_{ji}(0)\right|.$$
With condition (159), this results in
$$\mathrm{\Delta }S(t)=\mathrm{ln}\left|\underset{i}{}M_{ii}(t)\right|=\underset{i}{}\mathrm{ln}|M_{ii}(t)|.$$
(164)
Taking the variational derivative of equation (152), we get the equation
$$\frac{d}{dt}M(t)=J(y,t)M(t)$$
(165)
for the multiplier matrix (158), where
$$J(y,t)=[J_{ij}(y,t)],J_{ij}(y,t)\frac{\delta v_i(y,t)}{\delta y_j(t)},$$
(166)
is the Jacobian matrix. Substituting the entropy variation (161) into Eq. (156), we get
$$p(\beta ,t)=\frac{_i|M_{ii}(\beta ,t)|^1}{Z(t)},$$
(167)
with
$$Z(t)=\underset{i}{}|M_{ii}(\beta ,t)|^1d\beta .$$
Expression (164) defines the probability distribution of patterns labelled by a multiparameter $`\beta `$. This expression naturally connects the notion of probability and the notion of stability. Really, the multipliers are smaller by modulus for more stable solutions and, respectively, patterns, for which the probability is higher.
Another form of the distribution (164) can be derived as follows. Introduce the matrix
$$L(t)=[L_{ij}(t)],L_{ij}(t)\mathrm{ln}|M_{ij}(t)|.$$
(168)
Then the entropy variation (161) becomes
$$\mathrm{\Delta }S(t)=\mathrm{Tr}L(t).$$
(169)
Since the trace of a matrix does not depend on its representations, we may perform intermediate transformations of Eq. (166) using one particular representation and returning at the end to the form independent of representations. To this end, let us consider a representation when the multiplier matrix is diagonal. Because of Eq. (162) with the initial condition (159), the matrix $`M`$ is diagonal if and only if the Jacobian matrix is also diagonal. Then from the evolution equation (162) it follows that
$$M_{ii}(t)=\mathrm{exp}\left\{_0^tJ_{ii}(y(t^{}),t^{})๐t^{}\right\}.$$
Hence
$$L_{ii}(t)=_0^t\mathrm{\Lambda }_i(t^{})๐t^{},\mathrm{\Lambda }_i(t)\mathrm{Re}J_{ii}(t),$$
from where
$$\mathrm{Tr}L(t)=_0^t\mathrm{\Lambda }(t^{})๐t^{},\mathrm{\Lambda }(t)\underset{i}{}\mathrm{\Lambda }_i(t).$$
We assume that the state (151) is formed of real functions, so that the velocity field in the evolution equation (152) is also real. Then the eigenvalues of the Jacobian matrix (163) are either real or, if complex, come in complex conjugate pairs. Therefore
$$\underset{i}{}\mathrm{Re}J_{ii}(y,t)=\underset{i}{}J_{ii}(y,t)=\mathrm{Tr}J(y,t).$$
For the entropy variation (166) we obtain
$$\mathrm{\Delta }S(t)=_0^t\mathrm{\Lambda }(t^{})๐t^{},$$
(170)
where
$$\mathrm{\Lambda }(t)=\mathrm{Tr}J(y,t)$$
(171)
is called the contraction rate. The latter is given by the form independent of representations of the Jacobian matrix (163). With the entropy variance (167), the probability distribution (156) becomes
$$p(\beta ,t)=\frac{1}{Z(t)}\mathrm{exp}\left\{_0^t\mathrm{\Lambda }(\beta ,t^{})๐t^{}\right\},$$
(172)
where the contraction rate is defined in Eq. (168) and
$$Z(t)=\mathrm{exp}\left\{_0^t\mathrm{\Lambda }(\beta ,t^{})๐t^{}\right\}๐\beta .$$
The most probable pattern at a time $`t`$ corresponds to the maximum of the distribution (169),
$$\mathrm{abs}\underset{\beta }{\mathrm{max}}p(\beta ,t)\beta (t).$$
(173)
One may also define the average pattern at $`t`$ as corresponding to
$$\overline{\beta }(t)\beta p(\beta ,t)๐\beta .$$
The most probable and average patterns, in general, do not coincide, although this may happen, especially with increasing time. To illustrate the latter, consider a particular case when the contraction rate $`\mathrm{\Lambda }(\beta ,t)=\mathrm{\Lambda }(\beta )`$ does not depend on time. Then, as $`t\mathrm{}`$, we have
$$Z(t)=e^{\mathrm{\Lambda }(\beta )t}๐\beta \sqrt{\frac{2\pi }{\mathrm{\Lambda }^{\prime \prime }(\beta _0)t}}\mathrm{exp}\left\{\mathrm{\Lambda }(\beta _0)t\right\},$$
where $`\beta _0`$ is the point of the minimum of $`\mathrm{\Lambda }(\beta )`$, so that
$$\frac{d}{d\beta }\mathrm{\Lambda }(\beta )=0,\mathrm{\Lambda }^{\prime \prime }(\beta )\frac{d^2}{d\beta ^2}\mathrm{\Lambda }(\beta )>0(\beta =\beta _0).$$
In the distribution
$$p(\beta ,t)\sqrt{\frac{\mathrm{\Lambda }^{\prime \prime }(\beta _0)}{2\pi }t}\mathrm{exp}\left\{[\mathrm{\Lambda }(\beta )\mathrm{\Lambda }(\beta _0)]t\right\}$$
one may expand $`\mathrm{\Lambda }(\beta )`$ near $`\beta =\beta _0`$, which gives
$$p(\beta ,t)\frac{1}{\sqrt{2\pi }\sigma (t)}\mathrm{exp}\left\{\frac{(\beta \beta _0)^2}{2\sigma ^2(t)}\right\},\sigma (t)\frac{1}{\sqrt{\mathrm{\Lambda }^{\prime \prime }(\beta _0)t}}.$$
From here one finds
$$\underset{t\mathrm{}}{lim}p(\beta ,t)=\delta (\beta \beta _0).$$
In this way, if differential equations describing a nonequilibrium process have several solutions corresponding to different spatial patterns, the latter can be characterized by the probability distribution (169), with the contraction rate (168). In the case when the multiplier matrix (158) can be calculated, one may use the expression (164) of the probability distribution. If all patterns correspond to stable solutions, it is sufficient to analyse only the beginning of the process of pattern formation. Then for the entropy variation (167) we may write
$$\mathrm{\Delta }S(\beta ,t)\mathrm{\Lambda }(\beta ,0)t(t0).$$
Consequently, the most probable pattern, defined by the maximum of the probability distribution (169), that is, by the minimum of the entropy variation (167), is now characterized by the minimum of the contraction rate $`\mathrm{\Lambda }(\beta ,0)`$ at the initial time.
## 10 Turbulent Photon Filamentation
Spatial structures can appear in radiating systems if the radiation wavelength is much shorter than the system characteristic sizes . For instance, electric field in laser cavities can exhibit a state which bears some analogy with a superfluid vortex . The MaxwellโBloch equations for slowly varying field amplitudes have been shown to be analogous to hydrodynamic equations for compressible viscous fluid . The Fresnel number for optical systems plays the role similar to the Reynolds number for fluids. In the same way as when increasing the Reynolds number, the fluid becomes turbulent, there can appear optical turbulence when increasing the Fresnel number.
Spatial structures emerge from an initially homogeneous state with a break of spaceโtranslational symmetry. For small Fresnel numbers $`F5`$, such structures correspond to the emptyโcavity GaussโLaguerre modes imposed by the cavity geometry. These transverse structures can be described by expanding fields over the modal GaussโLaguerre functions \[87โ92\], which results in reasonable agreement with experiments for CO<sub>2</sub> and Na<sub>2</sub> lasers. For large Fresnel numbers $`F>10`$, the appearing structures are very different from those associated with emptyโcavity modes. The modal expansion is no longer relevant at large $`F`$, and the boundary conditions have little or no importance. The laser medium looks like divided in a large amount of parallel independently oscillating uncorrelated filaments \[93โ100\] the number of filaments being proportional to $`F`$, contrary to the case of small Fresnel numbers when the number of bright spots is proportional to $`F^2`$. This filamentation was observed in Dye and CO<sub>2</sub> lasers, as well as in other resonance media, even without resonators \[101โ105\]. The same type of patterns arises in active nonlinear media, such as photorefractive Bi<sub>12</sub>SiO<sub>20</sub> crystal pumped by a laser \[106โ109\]. In the latter media there are also two types of pattern formation: for small Fresnel numbers, the symmetry is imposed through the boundary, while for large Fresnel numbers, the symmetry is imposed by the bulk parameters. In the case of large $`F`$, there occurs a kind of selfโorganization with spontaneous spatial symmetry breaking . It is possible to easily notice a qualitative transition in the behaviour of photorefractive media as well as in that of lasers: In lowโ$`F`$ regime there are a few modes of regular arrangement of bright spots corresponding to the peaks of the GaussโLaguerre functions in cylindrical geometry, the number of modes being proportional to $`F^2`$. And in the highโ$`F`$ regime there are many modes spatially uncorrelated with each other, which is typical for spatiotemporal chaos, the number of the chaotic filaments being proportional to $`F`$. Shortโrange spatial correlation is characteristic for turbulence, this is why one calls the similar phenomenon in optics the optical turbulence.
The theory of selfโorganized photon filamentation in highโFresnelโnumber resonant media was suggested in Refs. \[33,111-116\], where the consideration was based on simplified models and only the stationary regime was analysed. The choice of filament radii was done by means of the variational principle, as is described in subsection 2.5. Here we present a more general and elaborate theory based on the evolution equations (68) to (70), which includes the description of temporal behaviour, and for defining the characteristics of filaments we employ the method of pattern selection developed in Sec. 9.
First, it is convenient to pass in Eqs. (68) to (70) to continuous representation replacing the sums by integrals according to the rule
$$\underset{i=1}{\overset{N}{}}=\rho (\stackrel{}{r})๐\stackrel{}{r},$$
where $`\rho (\stackrel{}{r})`$ is the spatial density of radiators. Wishing to return to the localized representation, one makes the replacement $`\rho (\stackrel{}{r})=_{i=1}^N\delta (\stackrel{}{r}\stackrel{}{r}_i)`$. In the case when the structure of matter is of no importance, it can be treated as uniform on average setting $`\rho (\stackrel{}{r})=\rho N/V`$. Cooperative optical phenomena are often considered in this representation of uniform medium . Let us stress that the uniformity of matter in no case requires the uniformity of fields or polarization. The solutions to Eqs. (68) to (70) can correspond to highly nonuniform structures.
Introduce the notation
$$f(\stackrel{}{r},t)f_0(\stackrel{}{r},t)+f_{rad}(\stackrel{}{r},t)$$
(174)
for an effective field acting on a radiator with the transition dipole $`\stackrel{}{d}`$. This field consists of the term
$$f_0(\stackrel{}{r},t)i\stackrel{}{d}\stackrel{}{E}_0(\stackrel{}{r},t)$$
(175)
due to an external electric field and of the term
$$f_{rad}(\stackrel{}{r},t)k_0<\stackrel{}{d}\stackrel{}{A}_{rad}(\stackrel{}{r},t)>$$
(176)
responsible for the action of other radiators. Taking into account Eq. (65), we have
$$f_{rad}(\stackrel{}{r},t)=\frac{3}{4}i\gamma \rho \left[\phi (\stackrel{}{r}\stackrel{}{r}^{})u(\stackrel{}{r}^{},t)\stackrel{}{e_d}^2\phi ^{}(\stackrel{}{r}\stackrel{}{r}^{})u^{}(\stackrel{}{r}^{},t)\right]๐\stackrel{}{r}^{},$$
(177)
where the continuous representation is used, and
$$\phi (\stackrel{}{r})\frac{e^{ik_0|\stackrel{}{r}|}}{k_0|\stackrel{}{r}|},\gamma \frac{4}{3}k_0^3d_0^2.$$
Then Eqs. (68) to (70) acquire the form
$$\frac{du}{dt}=(i\omega _0+\gamma _2)u+sf,\frac{ds}{dt}=2(u^{}f+f^{}u)\gamma _1(s\zeta ),$$
$$\frac{d|u|^2}{dt}=2\gamma _2|u|^2+s(u^{}f+f^{}u).$$
(178)
Notice that from the latter two equations one has
$$\frac{d}{dt}\left(s^2+4|u|^2\right)=2\gamma _1s(s\zeta )8\gamma _2|u|^2.$$
We consider a sample of the cylindrical shape typical of lasers. The seed laser field defining the cylinder axis is given by the sum of two running waves,
$$\stackrel{}{E}_0(\stackrel{}{r},t)=\stackrel{}{E}_1e^{i(kz\omega t)}+\stackrel{}{E}_1^{}e^{i(kz\omega t)},$$
(179)
which selects a longitudinal mode. The radius, $`R`$, and length, $`L`$, of the cylinder are such that the following inequalities are valid:
$$\frac{a}{\lambda }1,\frac{\lambda }{R}1,\frac{R}{L}1,$$
(180)
where $`a`$ is the mean distance between radiators and $`\lambda `$, wavelength. There are also the standard small parameters
$$\frac{\gamma _1}{\omega _0}1,\frac{\gamma _2}{\omega _0}1,\frac{|\mathrm{\Delta }|}{\omega _0}1,$$
(181)
with $`\mathrm{\Delta }\omega \omega _0`$ being detuning.
The solutions to Eqs. (175) are not necessarily uniform in the whole volume $`V=\pi R^2L`$ of the sample, but may have noticeable values only inside narrow regions of filamentary form, while being almost zero outside these filaments. Consider one such filament, and let us surround it by a cylinder of radius $`b`$ so that the magnitude of solutions is an order smaller at the surface of this enveloping cylinder than at its axis. If the profile of a filament is close to the Gaussian $`\mathrm{exp}(r^2/2r_f^2)`$, with $`r_f`$ being the filament radius, then
$$b=\sqrt{2\mathrm{ln}10}r_f.$$
(182)
In what follows we assume this relation between the radius $`b`$ of an enveloping cylinder and the radius $`r_f`$ of a filament.
Suppose that there are $`N_f`$ filaments in the volume of the sample, the axis of each filament being centered at a point $`\{x_n,y_n\}`$, with $`n=1,2,\mathrm{},N_f`$. Let us present the solutions to Eqs. (175) as expansions over enveloping cylinders,
$$u(\stackrel{}{r},t)=\underset{n=1}{\overset{N_f}{}}u_n(\stackrel{}{r},t)\mathrm{\Theta }_n(x,y)e^{ikz},s(\stackrel{}{r},t)=\underset{n=1}{\overset{N_f}{}}s_n(\stackrel{}{r},t)\mathrm{\Theta }_n(x,y),$$
(183)
where
$$\mathrm{\Theta }_n(x,y)\mathrm{\Theta }\left(b\sqrt{(xx_n)^2+(yy_n)^2}\right)$$
is a unitโstep function. The filaments are located randomly in the crossโsection of the sample, but so that their enveloping cylinders do not intersect with each other. The interaction between filaments is small, which follows from Eq. (174). This is why they do not form a regular lattice but are distributed randomly.
The function $`\phi (\stackrel{}{r})`$ in Eq. (174) oscillates at the distance $`\lambda `$, and the solutions $`u_n`$ and $`s_n`$ essentially change in the radial direction in the interval $`b`$. Assuming that
$$\frac{\lambda }{b}1,$$
(184)
we may say that, in the radial direction, the function $`\phi (\stackrel{}{r})`$ is fastly varying in space, as compared to the slow variation of $`u_n`$ and $`s_n`$. For the latter, we define the averages
$$u(t)\frac{1}{V_n}_{๐_n}u_n(\stackrel{}{r},t)๐\stackrel{}{r},s(t)\frac{1}{V_n}_{๐_n}s_n(\stackrel{}{r},t)๐\stackrel{}{r}$$
(185)
over the corresponding enveloping cylinder of the volume $`V_n\pi b^2L`$, where in the leftโhand side of Eq. (182) we, for short, do not write the index $`n`$.
The seed field (176) is needed mainly for selecting a longitudinal mode with cylindrical symmetry, but the amplitude of this field is small, so that
$$\frac{|\stackrel{}{d}\stackrel{}{E}_1|}{\gamma _2}1.$$
(186)
The excitation of radiators is accomplished by means of pumping characterized by the pumping parameter $`\zeta `$ in Eqs. (175).
Defining the effective coupling parameters
$$g\frac{3\gamma \rho }{4\gamma _2V_n}_{๐_n}\frac{\mathrm{sin}[k_0|\stackrel{}{r}\stackrel{}{r}^{}|k(zz^{})]}{k_0|\stackrel{}{r}\stackrel{}{r}^{}|}๐\stackrel{}{r}๐\stackrel{}{r}^{},$$
(187)
$$g^{}\frac{3\gamma \rho }{4\gamma _2V_n}_{๐_n}\frac{\mathrm{cos}[k_0|\stackrel{}{r}\stackrel{}{r}^{}|k(zz^{})]}{k_0|\stackrel{}{r}\stackrel{}{r}^{}|}๐\stackrel{}{r}๐\stackrel{}{r}^{},$$
(188)
and the collective frequency and width, respectively,
$$\mathrm{\Omega }\omega _0+g^{}\gamma _2s,\mathrm{\Gamma }\gamma _2(1gs),$$
(189)
for functions (182) we obtain the equations
$$\frac{du}{dt}=(i\mathrm{\Omega }+\mathrm{\Gamma })uis\stackrel{}{d}\stackrel{}{E}_1e^{i\omega t},$$
$$\frac{ds}{dt}=4g\gamma _2|u|^2\gamma _1(s\zeta )4\mathrm{I}\mathrm{m}\left(u^{}\stackrel{}{d}\stackrel{}{E}_1e^{i\omega t}\right),$$
(190)
$$\frac{d|u|^2}{dt}=2\mathrm{\Gamma }|u|^2+2s\mathrm{Im}\left(u^{}\stackrel{}{d}\stackrel{}{E}_1e^{i\omega t}\right).$$
Because of the inequalities (178) and (183), the solution $`u`$ in Eqs. (187) is fast, while $`s`$ and $`|u|^2`$ are slow in time. Using the scale separation approach, we find
$$u(t)=u_0e^{(i\mathrm{\Omega }+\mathrm{\Gamma })t}+\frac{s\stackrel{}{d}\stackrel{}{E}_1}{\omega \mathrm{\Omega }+i\mathrm{\Gamma }}\left[e^{i\omega t}e^{(i\mathrm{\Omega }+\mathrm{\Gamma })t}\right].$$
(191)
Introduce the parameter
$$\alpha \underset{\tau \mathrm{}}{lim}\frac{\mathrm{Im}}{\tau \mathrm{\Gamma }s}_0^\tau u^{}(t)\stackrel{}{d}\stackrel{}{E}_1e^{i\omega t}๐t,$$
(192)
characterizing the coupling of radiators with the seed field. This, with Eq. (188), gives
$$\alpha =\frac{|\stackrel{}{d}\stackrel{}{E}_1|^2}{(\omega \mathrm{\Omega })^2+\mathrm{\Gamma }^2}.$$
(193)
The latter, according to inequality (183), is small,
$$|\alpha |1.$$
(194)
Finally, defining the function
$$w|u|^2\alpha s^2,$$
(195)
we obtain the equations
$$\frac{ds}{dt}=4g\gamma _2w\gamma _1(s\zeta ),\frac{dw}{dt}=2\gamma _2(1gs)w.$$
(196)
The behaviour of solutions to Eqs. (193) essentially depends on the values of the coupling parameters (184) and (185). To evaluate the latter, we may notice that their integrands diminish and fastly oscillate at the distance of the wavelength $`\lambda `$. If condition (181) holds, we may neglect boundary effects in the integrals (184) and (185) writing approximately
$$_{๐_n}f(\stackrel{}{r}\stackrel{}{r}^{})๐\stackrel{}{r}๐\stackrel{}{r}^{}V_n_{๐_n}f(\stackrel{}{r})๐\stackrel{}{r}.$$
Then parameter (184) reduces to
$$g=\frac{3\pi \gamma \rho }{2\gamma _2}_0^br๐r_{L/2}^{L/2}\frac{\mathrm{sin}(k_0\sqrt{r^2+z^2}kz)}{k_0\sqrt{r^2+z^2}}๐z,$$
where $`r`$ is the radial variable. Because of the quasiresonance condition $`|\mathrm{\Delta }|\omega _0`$, we have $`k_0k`$. With the change of the variable $`xk(\sqrt{r^2+z^2}z)`$, we get
$$g=\frac{3\pi \gamma \rho }{2\gamma _2k}_0^br๐r_{kr^2/L}^{kL}\frac{\mathrm{sin}x}{x}๐x.$$
In this expression, one can replace $`kL\mathrm{}`$, thus obtaining
$$g=\frac{3\pi \gamma \rho }{2\gamma _2k}_0^b\left[\frac{\pi }{2}\mathrm{Si}\left(\frac{kr^2}{L}\right)\right]r๐r,$$
where the integral sine appears,
$$\mathrm{Si}(x)_0^x\frac{\mathrm{sin}t}{t}๐t=\frac{\pi }{2}+\mathrm{si}(x),\mathrm{si}(x)_{\mathrm{}}^x\frac{\mathrm{sin}t}{t}๐t.$$
Introducing the dimensionless quantity
$$\beta \frac{kb^2}{L}=\frac{2\pi b^2}{\lambda L},$$
(197)
we come to the coupling parameter
$$g=g(\beta )=\frac{3\pi \gamma \rho L}{4\gamma _2k^2}_0^\beta \left[\frac{\pi }{2}\mathrm{Si}(x)\right]๐x.$$
(198)
This can be integrated explicitly by means of the property
$$\mathrm{Si}(x)๐x=x\mathrm{Si}(x)+\mathrm{cos}x,$$
which results in
$$g(\beta )=\frac{3\pi \gamma \rho L}{4\gamma _2k^2}\left\{\beta \left[\frac{\pi }{2}\mathrm{Si}(\beta )\right]+1\mathrm{cos}\beta \right\}.$$
(199)
For the coupling parameter (185), one similarly finds
$$g^{}=g^{}(\beta )=\frac{3\pi \gamma \rho L}{4\gamma _2k^2}_0^\beta \mathrm{Ci}(x)๐x,$$
(200)
where the integral cosine occurs,
$$\mathrm{Ci}(x)_{\mathrm{}}^x\frac{\mathrm{cos}t}{t}๐t.$$
Integrating
$$\mathrm{Ci}(x)๐x=x\mathrm{Ci}(x)\mathrm{sin}x,$$
we finally get
$$g^{}(\beta )=\frac{3\pi \gamma \rho L}{4\gamma _2k^2}\left[\mathrm{sin}\beta \beta \mathrm{Ci}(\beta )\right].$$
(201)
To better understand the properties of the coupling parameters, we consider two limiting cases. When $`x1`$, then
$$\mathrm{Si}(x)x\frac{x^3}{18},\mathrm{Ci}(x)\gamma _E+\mathrm{ln}x\frac{x^2}{4},$$
where $`\gamma _E=0.577216`$ being the Euler constant. From here
$$g(x)\frac{3\pi \gamma \rho L}{4\gamma _2k^2}\left(\frac{\pi }{2}x\frac{1}{2}x^2\right),g^{}(x)\frac{3\pi \gamma \rho L}{4\gamma _2k^2}x|\mathrm{ln}x|.$$
In the opposite case, when $`x1`$, using
$$\mathrm{Si}(x)\frac{\pi }{2}\frac{\mathrm{cos}x}{x}\frac{\mathrm{sin}x}{x^2},\mathrm{Ci}(x)\frac{\mathrm{sin}x}{x}\frac{\mathrm{cos}x}{x^2},$$
we find
$$g(x)\frac{3\pi \gamma \rho L}{4\gamma _2k^2}\left(1+\frac{\mathrm{sin}x}{x}\right),g^{}(x)\frac{3\pi \gamma \rho L}{4\gamma _2k^2}\left(\frac{\mathrm{cos}x}{x}\right).$$
These asymptotic expressions help to analyse the dependence of the coupling parameters on the variable (194) changing in the interval
$$0<\beta 2F\left(F\frac{\pi R^2}{\lambda L}\right).$$
(202)
The stability analysis of Eqs. (193), similarly to that given in Ref. , shows that, for $`g\zeta <1`$, the solutions tend to the stationary stable point $`s_1^{}=\zeta ,w_1^{}=0`$, while for $`g\zeta >1`$, the stable fixed point is
$$s_2^{}=\frac{1}{g},w_2^{}=\frac{\gamma _1(g\zeta 1)}{4g^2\gamma _2}.$$
In this way, for all $`\beta `$ from the interval (199), except the sole case when $`g\zeta =1`$, there exists a stable fixed point, that is, almost all solutions are stable, independently of the value of $`\beta `$. Following the method of pattern selection from Sec. 9, we can equip the solutions labelled by $`\beta `$ with the probabilistic weights (169). The most probable, among all stable solutions, is that providing the minimum of the initial contraction rate, which for this case is
$$\mathrm{\Lambda }(\beta ,0)=\gamma _12\gamma _2(1gs_0).$$
(203)
The minimum of this rate requires that
$$\frac{dg}{d\beta }=0,s_0\frac{d^2g}{d\beta ^2}>0.$$
(204)
For $`s_0>0`$, one needs the minimum of $`g`$, which gives $`\beta =4.9`$. From Eq. (194), one has $`b=0.88\sqrt{\lambda L}`$. And the relation (179) yields
$$r_f=0.41\sqrt{\lambda L}(s_0>0).$$
(205)
When $`s_0<0`$, conditions (201) imply the maximum of $`g`$, for which $`\beta =1.92,b=0.55\sqrt{\lambda L}`$, and the filament radius is
$$r_f=0.26\sqrt{\lambda L}(s_0<0).$$
(206)
This is practically the same value as found for the filaments radius in Refs. \[33,111โ115\] by using the variational principle of subsection 2.5. When the system of radiators is not inverted at the initial time and becomes excited by means of a pulse characterized by the pumping parameter $`\zeta `$, one has to consider the filament radius (203) as corresponding to the most probable pattern. The number of filaments can be defined from the normalization condition
$$\frac{1}{V}s(\stackrel{}{r},t)๐\stackrel{}{r}=\zeta ,$$
(207)
assuming that the population difference equals $`+1`$ inside each filament of radius $`r_f`$ and $`1`$ outside of the filaments. Then the number of filaments is
$$N_f=\frac{1}{2}(1+\zeta )\left(\frac{R}{r_f}\right)^2.$$
(208)
The most probable filament radius (203) and the number of filaments (205) are in good agreement with the values observed in experiments \[93โ99,101โ105\]. The considered phenomenon of filamentation can be termed turbulent since the filaments are chaotically distributed in space and for sufficiently strong pumping, when $`g\zeta >1+\gamma _1/8\gamma _2`$, each filament is aperiodically flashing in time. The turbulent photon filamentation is a selfโorganized phenomenon due to the bulk properties of interacting radiators. It practically does not depend on boundary conditions and exists in both types of lasers, the resonatorโcavity lasers, such as CO<sub>2</sub> and Dye lasers \[93โ99\], as well as in the resonatorless dischargeโtube lasers, such as lasers on Ne, Tl, Pb, N<sub>2</sub>, and N$`{}_{}{}^{+}{}_{2}{}^{}`$ vapors \[101โ105\]. The turbulent filamentation is also principally nonlinear phenomenon. Thus, in lowโFresnelโnumber lasers ($`F5`$) the number of light spots is proportional to $`F^2`$. The same dependence of the number of coherent rays on $`F`$ is typical of the initial linearized stage of superfluorescence . However, for highโ$`F`$ lasers ($`F10`$) the number of filaments is proportional to $`F`$, which is in agreement with formula (205) giving $`N_fF`$.
## 11 Superradiant Spin Relaxation
When the initial state of a spin system is strongly nonequilibrium, different kinds of spin relaxation can occur. If there are no transverse external fields acting on spins, they relax to an equilibrium state by an exponential law with a longitudinal relaxation time $`T_1`$. When the motion of spins is triggered by a transverse magnetic field, the relaxation is again exponential but with a transverse relaxation time $`T_2`$ that is usually much shorter than $`T_1`$. A rather special relaxation regime arises, if the spin system is coupled to a resonator. This can be done by inserting the sample into a coil connected with a resonance electric circuit. Because of the action of resonator feedback field, the motion of spins can become highly coherent resulting in their ultrafast relaxation during a characteristic collective relaxation time much shorter than $`T_2`$ . This latter type of collective spin relaxation from a strongly nonequilibrium state in the presence of coupling with a resonator is the most difficult to realize experimentally and to describe theoretically. Experimental difficulties have been overcome in a series of observations of this phenomenon for a system of nuclear spins inside different paramagnetic materials \[120โ127\]. The collective relaxation time of this ultrafast coherent process is inversely proportional to the number of spins, $`N`$, and the intensity of magnetodipole radiation is proportional to $`N^2`$, in the same way as cooperative radiation time and radiation intensity of $`N`$ resonant atoms depend on this number in optic superradiance . This is why the process of collective coherent relaxation of spins has been called superradiant spin relaxation or, for short, spin superradiance. In the case of spin systems, what is usually measured is not the radiation intensity itself, which is rather weak, but the power of current induced in the resonant circuit . The enhancement of generated pulses by using resonators is, actually, well known in laser optics and is important for realizing superradiance of Rydberg atoms and recombination superradiance in electronโhole or electronโpositron plasmas . Resonators can be employed for modifying radiated pulses in optical superradiance . Note also the usage of resonators for amplifying the nuclear spin echo signals in magnets .
The appearance of strong correlations between spins is due to the resonator feedback field, but not to the photon exchange as it happens for atomic systems. Hence, various quantum effects existing in the interaction of electromagnetic field with atoms \[32,134โ137\] seem to be absent in the case of spin systems. Therefore it looked natural to try, for the theoretical description of relaxation in a spin system coupled with a resonator, to invoke the classical Bloch equations complimented by the Kirchhoff equation for the resonant electric circuit \[1,119,138โ140\]. However, these equations can provide a description of coherent spin relaxation only when the latter is triggered by a coherent pulse, similarly to the semiclassical BlochโMaxwell equations in optics . The phenomenon of the selfโorganized coherent spin relaxation cannot be described by the BlochโKirchhoff equations. Then, what initiates spin motion leading to the appearance of purely selfโorganized spin superradiance? This problem of the origin of pure spin superradiance was posed by Bloembergen and Pound . They also noticed that the thermal Nyquist noise of resonator cannot be a mechanism triggering the motion of spins, since the thermal relaxation time is proportional to the number of spins in the sample and, thus, the thermal damping is to be negligibly small for macroscopic samples. Nevertheless, this notice was forgotten by the following researchers who assumed that it is just the thermal noise of resonator which triggers the spin motion.
To resolve this controversy and to discover the genuine mechanisms originating the spin motion, it was necessary to turn to microscopic models. The system of nuclear spins is characterized by the Hamiltonian
$$\widehat{H}=\frac{1}{2}\underset{ij}{}H_{ij}\mu _n\underset{i}{}\stackrel{}{B}\stackrel{}{I}_i,$$
(209)
in which spins interact through the dipole potential
$$H_{ij}=\frac{\mu _n^2}{r_{ij}^3}\left[\stackrel{}{I}_i\stackrel{}{I}_j3\left(\stackrel{}{I}_i\stackrel{}{n}_{ij}\right)\left(\stackrel{}{I}_j\stackrel{}{n}_{ij}\right)\right],$$
where $`\mu _n`$ is the nuclear magnetic moment, $`\stackrel{}{I}_i`$ is a nuclear spin operator, $`r_{ij}=|\stackrel{}{r}_{ij}|,\stackrel{}{r}_{ij}=\stackrel{}{r}_i\stackrel{}{r}_j,\stackrel{}{n}_{ij}=\stackrel{}{r}_{ij}/r_{ij}`$. The total magnetic field
$$\stackrel{}{B}=H_0\stackrel{}{e}_z+H\stackrel{}{e}_x$$
contains an external magnetic field $`H_0`$ and a resonator feedback field $`H`$ defined by the Kirchhoff equation.
The temporal behaviour of a finite number of spins, with $`27N343`$, was analysed numerically by computer simulations \[144-149\]. From various cases studied, we present here some that give the general qualitative understanding of the whole picture. In Figs. 1-4, $`K_{coh}P_{coh}/P_{inc}`$ is a coherence coefficient, being the ratio of the coherent part of the current power $`P`$ to its incoherent part, and $`p_z`$ is the negative spin polarization. In Figs. 5โ11, $`C_{coh}I_{coh}/I_{inc}`$ is the coherence coefficient of the average magnetodipole radiation defined as in Eq. (90), with respect to the total radiation intensity $`I`$. The current power and radiation intensity are given in dimensionless units and time is measured in units of $`T_2`$. In the figure captions, $`p_z(0)`$ and $`p_x(0)`$ mean the corresponding polarization components at the initial time, $`\omega _0`$ is the Zeeman frequency, $`\omega `$ is the natural frequency of the resonant electric circuit and also a frequency of an alternating magnetic field, if any, the amplitude of the latter being denoted by $`h_0`$. The quantity
$$g\pi ^2\eta \frac{\rho _n\mu _n^2\omega _0}{\mathrm{}\mathrm{\Gamma }_2\omega }$$
(210)
is the effective coupling parameter, in which $`\eta `$ is a filling factor; $`\rho _n`$, nuclear density; and $`\mathrm{\Gamma }_2=T_2^1`$ is a line width. Computer simulations proved that pure spin superradiance does exist with no thermal noise involved.
However, computer simulations can provide only a qualitative picture, as the number of spins considered in such simulations is incomparably smaller than what one has in real macroscopic samples. Moreover, these simulations give no analytical formulas, making it difficult, if possible, to classify all relaxation regimes occurring when varying the numerous parameters of the system. Simplified models can also provide only a qualitative understanding.
An analytical solution of the evolution equations corresponding to the microscopic Hamiltonian (206) and a complete analysis of different relaxation regimes of nonequilibrium nuclear magnets coupled with a resonator has been done \[25,26,151-158\] by employing the scale separation approach. The evolution equations are written for the averages
$$u\frac{1}{N}\underset{i}{}<S_i^{}>,s\frac{1}{N}\underset{i}{}<S_i^z>,$$
(211)
where $`S_i^{}=S_i^xiS_i^y`$. Presenting local fluctuating fields through stochastic variables $`\xi _0`$ and $`\xi `$, one comes to the evolution equations
$$\frac{du}{dt}=i\left(\omega _0\xi _0+i\mathrm{\Gamma }_2\right)ui(\gamma _3h+\xi )s,$$
$$\frac{ds}{dt}=\frac{i}{2}\left(\gamma _3h+\xi \right)u^{}\frac{i}{2}\left(\gamma _3h+\xi ^{}\right)u\mathrm{\Gamma }_1(s\zeta ),$$
(212)
$$\frac{d}{dt}|u|^2=2\mathrm{\Gamma }_2|u|^2i\left(\gamma _3h+\xi \right)su^{}+i\left(\gamma _3h+\xi ^{}\right)su,$$
in which the resonator feedback field, $`h`$, in dimensionless units, satisfies the Kirchhoff equation
$$\frac{dh}{dt}+2\gamma _3h+\omega ^2_0^th(t^{})๐t^{}=2\kappa \frac{d}{dt}\left(u^{}+u\right)+\gamma _3f,$$
(213)
in which $`f`$ is an electromotive force, $`\gamma _3`$ is the resonator ringing width, and $`\kappa \pi \eta \rho _n\mu _n^2/\mathrm{}\gamma _3`$. The random local fields are defined as Gaussian stochastic variables with the stochastic averages
$$\xi _0^2=|\xi |^2=\mathrm{\Gamma }_{}^2,$$
(214)
where $`\mathrm{\Gamma }_{}`$ is the inhomogeneous dipole broadening. Because of the existence of the small parameters
$$\frac{\mathrm{\Gamma }_1}{\omega _0}1,\frac{\mathrm{\Gamma }_2}{\omega _0}1,\frac{\mathrm{\Gamma }_{}}{\omega _0}1,\frac{\gamma _3}{\omega _0}1,\frac{|\mathrm{\Delta }|}{\omega _0}1,$$
(215)
where $`\mathrm{\Delta }\omega \omega _0`$, the functions $`u`$ and $`h`$ can be classified as fast while $`s`$ and $`|u|^2`$ as slow.
Solving Eqs. (209) and (210), it was shown that the role of the thermal Nyquist noise in starting the relaxation process is negligible. But the main cause triggering the motion of spins, leading to coherent selfโorganization, is the action of nonsecular dipole interactions. This gives the answer to the question posed by Bloembergen and Pound : what is the origin of selfโorganized coherent relaxation in spin systems? All possible regimes of nonlinear spin dynamics have been analysed. When the nonresonant external pumping is absent, that is $`\zeta >0`$, there are seven qualitatively different transient relaxation regimes: free induction, collective induction, free relaxation, collective relaxation, weak superradiance, pure superradiance, and triggered superradiance .
In the presence of pumping, realized e.g. by means of dynamical nuclear polarization directing nuclear spins against an external constant magnetic field, one has $`\zeta 0`$. Then, as was shown using phenomenological equations , two stationary solutions can appear. In our approach, the behaviour of the system is as follows . When $`\zeta 0`$, three dynamical regimes can be observed, depending on the value of $`\zeta `$ with respect to the pump thresholds
$$\zeta _1=\frac{1}{g},\zeta _2=\frac{1}{g}\left(1+\frac{\mathrm{\Gamma }_1}{8\mathrm{\Gamma }_2}\right).$$
(216)
Analysing the equations for the slow variables $`s`$ and $`w`$, where
$$w|u|^2\frac{\mathrm{\Gamma }_{}^2}{\omega _0^2}s^2,$$
(217)
we find two fixed points
$$s_1^{}=\zeta ,w_1^{}=0;s_2^{}=\frac{1}{g},w_2^{}=\frac{\mathrm{\Gamma }_1(1+g\zeta )}{\mathrm{\Gamma }_2g^2}.$$
(218)
When $`\zeta _1<\zeta 0`$, the first fixed point is a stable node and the second one is a saddle point. For $`\zeta =\zeta _1`$, both points merge together, being neutrally stable. After the bifurcation at $`\zeta =\zeta _1`$, in the region $`\zeta _2\zeta <\zeta _1`$, the first fixed point looses its stability becoming a saddle point while the second fixed point becomes a stable node. Finally, when $`\zeta <\zeta _2`$, the second fixed point turns into a stable focus, and the first one continues to be a saddle point. In this way, there are three qualitatively different lasting relaxation regimes induced by the pumping . The first one is the incoherent monotonic relaxation to the first stationary solution $`s_1^{},w_1^{}`$. The second regime is the coherent monotonic relaxation to the second stationary solution $`s_2^{},w_2^{}`$, although the level of coherence may be rather low. And the third case is the coherent pulsing relaxation to the second fixed point. This unusual regime of pulsing relaxation was observed experimentally . Here we present the results of numerical solution of the evolution equations for the slow variables $`s=z(t)`$ and $`w(t)`$ defined in Eq. (214). Different cases of the pulsing regime are clearly demonstrated in Figs. 12 to 18. In the corresponding figure captions we use the notation $`z_0=z(0),w_0=w(0)`$, and $`\gamma \gamma _1/\gamma _2`$. Everywhere in Figs. 12 to 17, the pump parameter is $`\zeta =0.5`$, and in Fig. 18 this parameter is varied. The coupling parameter (207) is always $`g=10`$.
The problem of superradiant spin relaxation can be generalized to the case of nuclei incorporated into a ferromagnetic matrix, where nuclear and electron spins interact through hyperline forces. Some model studies of this case have been undertaken \[160โ162\], and a general microscopic theory has also been developed . The latter theory makes it possible to discover all feasible causes triggering the process of selfโorganized coherent relaxation. The most important such causes are the dipole hyperfine interactions, dipole nuclear interactions, and the transverse magnetocrystalline anisotropy.
## 12 Negative Electric Current
The study of electric processes in semiconductors is important for describing and modelling semiconductor devices . One of the most difficult problems is the consideration of strongly nonequilibrium phenomena in essentially nonuniform semiconductors. Nonequilibrium and nonuniform distributions of charge carriers can be formed in several ways, for instance, by means of external irradiation . Transport properties of semiconductors with essentially nonuniform distribution of charge carriers can be rather specific. For example, in a sample, biased with an external constant voltage, the resulting electric current may turn against the latter displaying the transient effect of negative electric current \[3,166โ168\].
Transport properties of semiconductors are usually described by the semiclassical driftโdiffusion equations . In what follows a plane device, of area $`A`$ and length $`L`$ is considered, which is biased with a constant voltage $`V_0`$. It is convenient to pass to dimensionless quantities, measuring the space variable $`x`$ in units of $`L`$, time in units of the transit time
$$\tau _0\frac{L^2}{\mu V_0},\mu \mathrm{min}\{|\mu _i|\},$$
where $`\mu _i`$ is a mobility of the $`i`$โtype carriers. And the characteristic quantities
$$\rho _0\frac{Q_0}{AL},Q_0\epsilon AE_0,E_0\frac{V_0}{L},$$
$$j_0\frac{Q_0}{A\tau _0},D_0\mu V_0,\xi _0\frac{\rho _0}{\tau _0},$$
are employed for measuring other physical values which are used below.
The driftโdiffusion equations consist of the continuity equations
$$\frac{\rho _i}{t}+\mu _i\frac{}{x}\left(\rho _iE\right)D_i\frac{^2\rho _i}{x^2}+\frac{\rho _i}{\tau _i}=\xi _i,$$
(219)
for each type of charge carriers, and of the Poisson equation
$$\frac{E}{x}=4\pi \underset{i}{}\rho _i$$
(220)
for the electric field $`E(x,t)`$. Here $`\rho _i(x,t)`$ is a charge density; $`\mu _i,D_i`$, and $`\tau _i`$ are mobility, diffusion coefficient, and relaxation time, respectively; $`\xi _i`$ is a generationโrecombination noise . The sample is biased with an external constant voltage, which in our dimensionless notation implies that
$$_0^1E(x,t)๐x=1.$$
(221)
At the initial time, the distribution of charge carriers
$$\rho _i(x,0)=f_i(x)$$
(222)
is assumed to be nonuniform.
The total electric current through the semiconductor sample is
$$J(t)_0^1j(x,t)๐x,$$
(223)
where the density of current
$$j=\underset{i}{}\left(\mu _iED_i\frac{}{x}\right)\rho _i+\frac{1}{4\pi }\frac{E}{t}.$$
(224)
Because of the voltage integral (218), one has
$$_0^1\frac{}{t}E(x,t)๐x=0.$$
(225)
It is also possible to show that
$$\underset{\tau \mathrm{}}{lim}\frac{1}{\tau }_0^\tau \frac{}{x}E(x,t)dt=0.$$
(226)
This means that the function $`E`$ can be considered as slow on average in time and in space. Then, treating $`E`$ as a quasiโinvariant, one may find the solutions to Eqs. (216) and (217) in order to analyse their general spaceโtime behaviour and to find conditions when the effect of negative electric current could arise. Such negative current can appear only when the initial charge distribution is essentially nonuniform. For example, if this initial charge distribution forms a narrow layer located at the point $`x=a`$, then the total current (220) becomes negative for a transient interval of time in the vicinity of $`t=0`$, if one of the following conditions holds true:
$$a<\frac{1}{2}\frac{1}{4\pi Q}\left(Q>\frac{1}{2\pi }\right),\mathrm{or}a>\frac{1}{2}+\frac{1}{4\pi |Q|}\left(Q<\frac{1}{2\pi }\right),$$
(227)
where
$$Q\underset{i}{}Q_i,Q_i_0^1\rho _i(x,0)๐x.$$
The effect of the negative electric current can be employed for various purposes, as is discussed in Refs. . For instance, when the initial charge layer is formed by an ion beam irradiating the semiconductor sample, the location $`a`$ corresponds to the ion mean free path. In this case, by measuring the negative current $`J(0)`$, one can define this mean free path
$$a=\frac{1}{2}\frac{1}{4\pi Q}\left[1\frac{J(0)}{_i\mu _iQ_i}\right].$$
(228)
This formula is valid for both positive and negative values of $`Q`$.
Equations (216) and (217) have also been solved numerically , which confirmed the appearance of the negative electric current. Two cases were analysed, with one layer of charge carriers and with two such layers. Here we present the results of calculations for the doubleโlayer case. The initial charge distributions (219) are given by the Gaussians
$$f_i(x)=\frac{Q_i}{Z_i}\mathrm{exp}\left\{\frac{(xa_i)^2}{2b_i}\right\},$$
in which $`0a_i1`$ and
$$Q_i=_0^1f_i(x)๐x,Z_i=_0^1\mathrm{exp}\left\{\frac{(xa_i)^2}{2b_i}\right\}๐x.$$
The positive charge carriers, with $`\mu _1=1`$ and $`Q_1=1`$, form the left layer centered at $`a_1=a`$, while the negative charge carriers form the layer centered at $`a_2=1a`$. We keep in mind the relation $`D_2=3D_1`$ for the diffusion coefficients, typical for holes and electrons, and we set $`D_1=10^3`$. For short, we use the notation $`\tau _1^1=\tau _2^1=\gamma `$ and $`b_1=b_2=b`$. The generationโrecombination noise is neglected, which is admissible at the initial stage of the process. As the boundary conditions, we accept the absence of diffusion through the semiconductor surface, which implies the Neumann boundary condition
$$\frac{}{x}\rho _i(x,t)=0(x=0,x=1).$$
In Figs. 19 to 24, we present the total current (220) as well as the electric current through the left surface, $`J(0,t)j(0,t)`$ and through the right surface, $`J(1,t)j(1,t)`$, defined by the current (221) at $`x=0`$ or $`x=1`$, respectively.
## 13 Magnetic Semiconfinement of Atoms
Dynamics of neutral atoms in nonuniform magnetic fields concerns problems of current experimental and theoretical interest. By means of such fields, atoms can be confined inside magnetic traps, which allows to accomplish various experiments with the systems of trapped atoms. Recently, BoseโEinstein condensation has been attained in a dilute gas of trapped atoms of <sup>87</sup>Rb , <sup>7</sup>Li , Na , and H$``$ . The details on theory and experiment can be found in reviews \[174โ176\]. The BoseโEinstein condensate is believed to form, at least partially, a coherent state. If it would be possible to construct a device emitting a coherent atomic beam, this would be analogous to a laser radiating a coherent photon ray. This is why one may call the device, emitting a coherent atomic beam, an atom laser \[177โ184\]. An output coupler, coherently extracting condensed atoms form a trap, was demonstrated recently \[185โ187\]. But in these demonstrations, the atoms, when escaping from a trap, fly out more or less in all directions, with anisotropy formed only by the gravitational force. While the very first condition on a laser is that its output is highly directional, with the possibility of varying the beam direction .
A mechanism for creating wellโcollimated beams of neutral atoms was advanced in Refs. \[188โ192\]. This mechanism suggests an output coupler that extracts trapped atoms in the form of a directed beam.
The motion of neutral atoms in magnetic fields can be described by the semiclassical equations for the quantumโmechanical average of the realโspace coordinate $`\stackrel{}{r}=\{r_\alpha \}`$, where $`\alpha =x,y,z`$, and for the average $`\stackrel{}{S}=\{S_\alpha \}`$ of the spin operator \[193โ195\]. The first equation writes
$$m\frac{d^2r_\alpha }{dt^2}=\mu _0\stackrel{}{S}\frac{\stackrel{}{B}}{r_\alpha }+mg_\alpha +f_\alpha ,$$
(229)
where $`m`$ and $`\mu _0`$ are mass and magnetic moment of an atom; $`\stackrel{}{B}`$ is a magnetic field; $`g_\alpha `$ is a component of the standard gravitational acceleration; and $`f_\alpha `$ is a collision force component. The equation for the average spin is
$$\mathrm{}\frac{d\stackrel{}{S}}{dt}=\mu _0\stackrel{}{S}\times \stackrel{}{B}.$$
(230)
The total magnetic field
$$\stackrel{}{B}=\stackrel{}{B}_1+\stackrel{}{B}_2,$$
$$\stackrel{}{B}_1=B_1^{}\left(x\stackrel{}{e}_x+y\stackrel{}{e}_y+\lambda z\stackrel{}{e}_z\right),\stackrel{}{B}_2=B_2\left(h_x\stackrel{}{e}_x+h_y\stackrel{}{e}_y\right),$$
(231)
where $`|\stackrel{}{h}|=1`$, consists of the quadrupole field $`\stackrel{}{B}_1`$, typical of quadrupole magnetic traps, and of a transverse field, e.g., of a rotating field . In the quadrupole field, $`\lambda `$ is the anisotropy parameter.
It is convenient to pass to the dimensionless space variable, measuring the components of $`\stackrel{}{r}`$ in units of the characteristic length
$$R_0\frac{B_2}{B_1^{}}.$$
(232)
Introduce the characteristic frequencies by the relations
$$\omega _1^2\frac{\mu _0B_1^{}}{mR_0},\omega _2\frac{\mu _0B_2}{\mathrm{}},\omega \underset{t}{\mathrm{max}}\left|\frac{d\stackrel{}{h}}{dt}\right|.$$
(233)
Also, we define
$$\delta _\alpha \frac{g_\alpha }{R_0\omega _1^2},\gamma \xi _\alpha \frac{f_\alpha }{mR_0},$$
(234)
where $`\gamma `$ is a collision rate and $`\xi _\alpha `$ can be treated as a random variable with the stochastic averages
$$\xi _\alpha (t)=0,\xi _\alpha (t)\xi _\beta (t^{})=2D_\alpha \delta _{\alpha \beta }\delta (tt^{}),$$
in which $`D_\alpha `$ is a diffusion rate. Then Eq. (226) can be written as the stochastic differential equation
$$\frac{d^2\stackrel{}{r}}{dt^2}=\omega _1^2\left(S_x\stackrel{}{e}_x+S_y\stackrel{}{e}_y+\lambda S_z\stackrel{}{e}_z+\stackrel{}{\delta }\right)+\gamma \stackrel{}{\xi },$$
(235)
and Eq. (227) acquires the form
$$\frac{d\stackrel{}{S}}{dt}=\omega _2\widehat{A}\stackrel{}{S},$$
(236)
in which the antisymmetric matrix $`\widehat{A}=[A_{\alpha \beta }]`$ has the elements
$$A_{\alpha \beta }=A_{\beta \alpha },A_{\alpha \alpha }=0,$$
$$A_{12}=\lambda z,A_{23}=x+h_x,A_{31}=y+h_y.$$
Assuming the occurrence of the small parameters
$$\left|\frac{\gamma }{\omega _1}\right|1,\left|\frac{\omega _1}{\omega _2}\right|1,\left|\frac{\omega }{\omega _2}\right|1,$$
(237)
we may classify the variables $`\stackrel{}{r}`$ and $`\stackrel{}{h}`$ as slow, compared to the fast spin variable $`\stackrel{}{S}`$. Then Eq. (233) can be solved yielding
$$\stackrel{}{S}(t)=\underset{i=1}{\overset{3}{}}a_i\stackrel{}{b}_i(t)\mathrm{exp}\{\beta _i(t)\},$$
(238)
where
$$a_i=\stackrel{}{S}(0)\stackrel{}{b}_i(0),$$
$$\stackrel{}{b}_i(t)=\frac{1}{\sqrt{C_i}}\left[\left(A_{12}A_{23}\alpha _iA_{31}\right)\stackrel{}{e}_x+\left(A_{12}A_{31}+\alpha _iA_{23}\right)\stackrel{}{e}_y+\left(A_{12}^2+\alpha _i^2\right)\stackrel{}{e}_z\right],$$
$$C_i=\left(A_{12}^2|\alpha _i|^2\right)^2+\left(A_{12}^2+|\alpha _i|^2\right)\left(A_{23}^2+A_{31}^2\right),$$
$$\alpha _{1,2}=\pm i\alpha ,\alpha _3=0,\alpha ^2A_{12}^2+A_{23}^2+A_{31}^2,\beta _i(t)=\omega _2_0^t\alpha _i(t^{})๐t^{}.$$
Substituting Eq. (235) into the rightโhand side of Eq. (232) and averaging the latter over time and over stochastic variables, we obtain
$$\frac{d^2\stackrel{}{r}}{dt^2}=\stackrel{}{F}+\omega _1^2\stackrel{}{\delta },$$
(239)
where
$$\stackrel{}{F}\omega _1^2a_3<b_3^x\stackrel{}{e}_x+b_3^y\stackrel{}{e}_y+\lambda b_3^z\stackrel{}{e}_z>,$$
$$a_3=\frac{(x+h_x^0)S_x^0+(y+h_y^0)S_y^0+\lambda zS_z^0}{[(x+h_x^0)^2+(y+h_y^0)^2+\lambda ^2z^2]^{1/2}},\stackrel{}{b}_3=\frac{(x+h_x)\stackrel{}{e}_x+(y+h_y)\stackrel{}{e}_y+\lambda z\stackrel{}{e}_z}{[(x+h_x)^2+(y+h_y)^2+\lambda ^2z^2]^{1/2}},$$
angle brackets imply time averaging and $`h_\alpha ^0h_\alpha (0),S_\alpha ^0S_\alpha (0)`$. For the rotating transverse field, with
$$h_x=\mathrm{cos}\omega t,h_y=\mathrm{sin}\omega t,$$
(240)
we find
$$\stackrel{}{F}=\frac{\omega _1^2[(1+x)S_x^0+yS_y^0+\lambda zS_z^0](x\stackrel{}{e}_x+y\stackrel{}{e}_y+2\lambda ^2z\stackrel{}{e}_z)}{2[(1+2x+x^2+y^2+\lambda ^2z^2)(1+x^2+y^2+\lambda ^2z^2)]^{1/2}}.$$
The motion of atoms, described by Eq. (236), essentially depends on the initial state, which, as is known , can be prepared in an arbitrary way. Suppose that atoms, after being laser cooled in a magnetoโoptical trap , are loaded into a magnetic trap where they are further cooled by evaporative cooling down to sufficiently low temperatures, so that there is a portion of atoms with low velocities, which are located close to the trap center. If the initial spin condition for these atoms is such that $`S_x^0<0`$ and $`S_y=S_z=0`$, then the atoms are confined inside the trap moving in an approximately harmonic potential. The gradient of the quadrupole field supplies the levitating force to support atoms against gravity. The combination of the magnetic field and gravity produces a very nearly harmonic confining potential within the trap volume in all three dimensions .
The semiconfining regime of motion \[188โ192\] can be realized by preparing for the spin variable nonadiabatic initial conditions
$$S_x^0=S_y^0=0,S_z^0S0.$$
(241)
Such conditions can be arranged in several ways. One possibility could be to confine atoms in a trap, where all atoms are polarized having their spins in the $`z`$ direction, as e.g. in the trap of Ref. , being a quadrupole trap with a bias field along the $`z`$ axis. Then the longitudinal bias field is quickly switched off, and at the same time, a transverse field is switched on, which would correspond to the sudden change of potential . Another way could be to prepare spin polarized atoms in one trap and quickly load them into another trap with the required field configuration. Atoms can be prepared practically $`100\%`$ polarized , with the spinโspin relaxation time reaching $`100`$ s . The possibility of realizing two ways of transferring atoms from one trap to another, by means of sudden transfer as opposed to adiabatic transfer, is discussed in Ref. . The third way of organizing the nonadiabatic initial conditions (238) could be by acting on the trapped atoms with a short pulse of strong magnetic field, polarizing atomic spins in the desired way.
With the initial conditions (238), the motion of atoms becomes axially restricted from one side, depending on the sign of $`\lambda S`$. Atoms fly out of the trap predominantly in one direction, forming a wellโcollimated beam \[188โ192\]. This mechanism can be used for atom lasers. Another possibility could be to study the dynamics of binary mixtures of Bose systems, where the effect of conical stratification can arise. The mixtures of two condensates have been realized for rubidium and sodium , and the dynamics of two rubidium condensates was observed in Ref. .
When solving equation (236) for the realistic case of a finite trap, one should take into account the trap shape factor, which can be written in the Gaussian form
$$\phi (\stackrel{}{r})=\mathrm{exp}\left(\frac{x^2+y^2}{R^2}\frac{z^2}{L^2}\right),$$
where $`R`$ and $`L`$ are the trap radius and length. The relation between the latter can be quite different for different traps, starting from almost spherical traps, where $`RL`$, to needleโshape traps, with $`R/L10^3`$, as for IoffeโPritchard magnetic traps . Accepting the initial spin conditions (238), and using the notation
$$f(\stackrel{}{r})\frac{\phi (\stackrel{}{r})}{[(1+2x+x^2+y^2+\lambda ^2z^2)(1+x^2+y^2+\lambda ^2z^2)]^{1/2}},$$
from Eq. (236) we obtain
$$\frac{d^2x}{dt^2}=\omega _1^2\left(\frac{\lambda }{2}Sfzx+\delta _x\right),\frac{d^2z}{dt^2}=\omega _1^2\left(\lambda ^3Sfz^2+\delta _z\right),$$
(242)
where the equation for $`y`$, being similar to that for $`x`$, is not written down. Note that instead of the Gaussian shape factor for the trap, one could opt for
$$\phi (\stackrel{}{r})=1\mathrm{\Theta }(xR)\mathrm{\Theta }(yR)\mathrm{\Theta }\left(|z|\frac{1}{2}L\right),$$
with $`\mathrm{\Theta }()`$ being the unitโstep function.
Equations (239) were analysed both analytically and numerically \[188โ192\]. Their solutions display the semiconfined regime of motion. Taking into account random pair collisions in Eq. (232) shows that atomic collisions do not disturb the semiconfined motion provided that temperature $`T`$ is sufficiently low, satisfying the condition
$$\frac{k_BT\mathrm{}\rho ^2a_s^2}{m^2\omega _1^3}1,$$
(243)
in which $`\rho `$ is the density of atoms and $`a_s`$, their scattering length. The semiconfined regime of motion makes it possible to form wellโcollimated beams on neutral atoms by means of only magnetic fields.
## 14 Nuclear Matter Lasing
The natural question that arises after talking about atom lasers is whether there can be produced matter waves corresponding to other Bose particles, which could be employed for lasing. One such possibility is related to the creation of large number of pions in hadronic, nuclear, and heavyโion collisions. If the density of pions appearing in the course of these collisions is sufficiently high, then correlations between pions can result in the formation of coherent state and in the feasibility of realizing a pion laser . Pions are not the sole type of Bose particles arising in nuclear matter under extreme conditions characteristic of fireballs produced in highโenergy collisions . There are plenty of reviews devoted to the state of nuclear matter at extreme conditions, including the region of deconfinement transition. Here we cite only some recent of such reviews \[214โ217\].
The very first necessary condition that is required for lasing is to be able to generate Bose particles with sufficiently high density. Therefore, in order to answer the question what kind of Bose particles appearing in nuclear matter under extreme conditions could be used for lasing, one has, first of all, to find out what are these Bose particles and under what conditions their density is maximal. In this section, we give a very brief account of an analysis based on the multichannel model of nuclear matter \[217โ221\]. The main idea in constructing this model goes back to the Weinberg approach for describing composite particles \[222โ224\], with effective Hamiltonians that are assumed to be a result of the FockโTani transformation . Now we shall not plunge into the details of the multichannel model, which can be found in Refs. , but we shall present some figures and will formulate the conclusion of an analysis with regard to the most probable candidates for nuclear matter lasing.
When rising temperature or density, nuclear matter exhibits a transition from hadron state to quarkโgluon state. This transition is often assumed to be a sharp firstโorder transition. Lattice numerical simulations for the quarkless $`SU(3)`$ gauge model show that deconfinement is really a firstโorder phase transition , which is in agreement with the multichannel model. Figures 25 to 27 illustrate the behaviour of some thermodynamic characteristics, normalized to the corresponding StefanโBoltzmann limits, for the case of the $`SU(3)`$ gluonโglueball mixture. Figure 28 shows the related glueball channel probability. The sharpness on the deconfinement transition essentially depends on the interactions between particles or on their radii, when the composite particles are treated as bags .
In the case of realistic nuclear matter, deconfinement is rather a gradual crossover but not a genuine phase transition . Then all thermodynamic characteristics change continuously, without jumps. This concerns as well the channel probabilities. Thus, in Figs. 29, 30 the channel probabilities of nucleons and dibaryons are shown as functions of baryon density normalized to the normal baryon density of nuclear matter $`n_{0B}=0.167`$ fm<sup>-3</sup>. The possible appearance of dibaryons is of special interest since they, being bosons, can form a Bose condensate \[217,228โ230\].
Summarizing the results of the analysis , three types of Bose particles can appear in nuclear matter in large quantities: pions, dibaryons, and gluons. The maximum of the pion channel probability, reaching $`w_\pi =0.6`$, occurs in the vicinity of the deconfinement transition at $`T160`$ MeV and low baryon densities $`n_B<n_{0B}`$. Dibaryons can appear mainly at low temperatures $`T<20`$ MeV and relatively high baryon densities $`n_B10n_{0B}`$, where their channel probability $`w_60.7`$. Large amount of gluons emerges only at high temperatures $`T>160`$ MeV. In addition, one should keep in mind that gluons cannot be observed as free particles.
Talking about possible pion, dibaryon, or gluon lasing from nuclear matter, we have touched here just one necessary condition, trying to find out when these Bose particles can appear in large quantities. To realize such a lasing in reality will, certainly, require to solve a number of other problems. But, anyway, to understand the conditions when this lasing could be plausible in principle is the necessary first step.
## 15 Conclusion
We have described a general method for treating strongly nonequilibrium processes in statistical systems. This method is called the Scale Separation Approach since its basic idea is to try to separate different characteristic scales of time and space variables. The idea itself is, of course, not new and we have employed some known techniques. What is original in our approach is: (i) The combination of several methods and their adjustment to the problems of nonequilibrium statistical mechanics. (ii) The generalization of the averaging method to stochastic and partial differential equations. (iii) Probabilistic solution of the problem of pattern selection.
The scale separation approach has been shown to be very useful for describing cooperative phenomena in the interaction of radiation with matter. To emphasize the generality of the approach, it is illustrated here by several different physical examples, whose common feature is that the related evolution equations are nonlinear differential or integroโdifferential stochastic equations. Such equations, as is known, are difficult to solve. The scale separation approach makes it possible to find accurate approximate solutions. The accuracy of these solutions has been confirmed by numerical calculations and by comparison with experiment, when available. Using this approach, several interesting physical problems have been solved and new effects are predicted. Among the most interesting applications we would like to emphasize the following.
Collective Liberation of Light happens when en ensemble of resonant atoms is doped into a medium with polariton band gap. If the transition frequency of an atom is inside this prohibited gap, then atomic spontaneous emission is strongly suppressed, which is termed localization of light. Although spontaneous emission of a single atom is prohibited, a collective of such atoms can radiate due to their coherent interactions. As a result of this coherent radiation, light becomes partially liberated. We have advanced dynamical theory of this light liberation for the realistic situation when the radiation wavelength is smaller than the linear sizes of the sample (see Sec. 6).
Mรถssbauer Magnetic Anomaly has puzzled researches for many years. This anomaly consists in a strong increase of the area under the Mรถssbauer spectrum, below the temperature of magnetic phase transition. Several explanations of this anomaly have been suggested. We have thoroughly analysed this phenomenon and concluded that previously suggested mechanisms cannot explain this anomaly but that its origin is rather in the increase of inhomogeneous broadening of Mรถssbauer nuclei, which is due to the arising magnetic field (see Sec. 8).
Turbulent Photon Filamentation in resonant media is an intriguing example of self-organization in a strongly nonequilibrium system, whose dynamical theory was absent. We have developed such a theory, based on the probabilistic approach to pattern selection, and showed that it gives agreement with experiment (see Sec. 10).
Superradiant Spin Relaxation occurs in a system of spins coherently interacting with each other through resonator feedback field. This ultrafast coherent relaxation is similar to superradiance in optical systems, because of which the term spin superradiance was coined. Contrary to its optical counterpart, the origin of purely self-organized spin superradiance has not been understood for about 40 years, after Bloembergen and Pound posed this problem in 1954. We have developed a theory of nonlinear spin dynamics, based on a microscopic Hamiltonian, elucidated the origin of pure spin superradiance, and described all main regimes of spin relaxation, without pumping as well as in the presence of the latter (see Sec. 11).
Negative Electric Current is a rather unusual effect, when electric current flows against an applied voltage. This is a transient effect that can occur in nonuniform semiconductors. We have predicted this effect and suggested its theory (see Sec. 12).
Magnetic Semiconfinement of Atoms is another effect we predict. This effect can serve as a mechanism for creating wellโcollimated beams of neutral atoms by means of magnetic fields. It can be used to form coherent beams of Bose atoms from atom lasers. We have presented a theory of this effect (see Sec. 13).
The possibility of treating nonequilibrium processes in nonlinear systems of quite different nature has become possible owing to the Scale Separation Approach, which provides accurate approximate solutions to complicated systems of differential and integroโdifferential equations.
Acknowledgement
We are grateful for discussions and useful advice to V.S. Bagnato, N.A. Bazhanov, C.M. Bowden, M.G. Cottam, V.I. Emelyanov, R. Friedberg, S.R. Hartmann, V.K. Henner, Vl.V. Kocharovsky, J.T. Manassah, A.N. Oraevsky, T. Ruskov, V.V. Samartsev, M.A. Singh, and R. Tanas. We appreciate the contribution of all our coauthors.
Figure Captions
Fig.1. Coherence coefficient $`K_{coh}`$, current power $`P`$, and spin polarization $`p_z`$ as functions of time for two different coupling parameters defined in Eq. (207), $`g_1`$ (solid line) and $`g_2`$ (dashed line), with the relation $`g_1/g_2=10`$.
Fig.2. The same as in Fig. 1 for two different Zeeman frequencies, $`\omega _{01}`$ (solid line) and $`\omega _{02}`$ (dashed line), related by the ratio $`\omega _{01}/\omega _{02}=5`$.
Fig.3. The same functions as in Fig. 1 for different initial polarizations, $`p_{z1}(0)`$ (solid line) and $`p_{z2}(0)`$ (dashed line), with the relation $`p_{z1}/p_{z2}(0)=2`$.
Fig.4. The same functions as in Fig. 1 for different initial transverse polarizations, $`p_{x1}(0)`$ (solid line) and $`p_{x2}(0)`$ (dashed line), with the relation $`p_{x1}/p_{x2}(0)=0.5`$.
Fig.5. Coherence coefficient $`C_{coh}`$, radiation intensivity $`I`$, and spin polarization $`p_z`$ versus time for $`p_z(0)=0.48`$ and different parameters: $`\omega _0=200,g=25`$ (solid line); $`\omega _0=40,g=25`$ (dashed line); and $`\omega _0=40,g=2.5`$ (solid line with crosses).
Fig.6. Coherence coefficient $`C_{coh}`$, radiation intensivity $`I`$, and spin polarization $`p_z`$ as functions of time in the case of switchedโoff resonatorโspin coupling ($`g=0`$). The varied parameters are: $`\omega _0=200,p_x(0)=0.48`$ (solid line); $`\omega _0=20,p_x(0)=0.48`$ (dashed line); and $`\omega _0=200,p_x(0)=0.20`$ (solid line with crosses).
Fig.7. The same as in Fig. 6 for $`p_x(0)=0.48`$ and for different Zeeman frequencies: $`\omega _0=1000`$ (solid line); $`\omega _0=200`$ (dashed line); $`\omega _0=50`$ (solid line with crosses); and $`\omega _0=200`$ with switchedโoff dipole interaction (solid line with triangles).
Fig.8. The same as in Fig. 6 for $`p_x(0)=0.48`$ but in the presence of an alternating magnetic field with the frequency $`\omega =\omega _0`$ and different amplitudes: $`h_{01}`$ (solid line); $`h_{02}`$ (dashed line); where $`h_{01}/h_{02}=10`$; and $`h_{03}=0`$ (solid line with crosses).
Fig.9. The same as in Fig. 8 but for $`p_x(0)=0.48`$ and different amplitudes of the alternating field: $`h_{01}`$ (solid line); $`h_{02}`$ (dashed line); and $`h_{03}`$ (solid line with crosses), where the amplitude relations are $`h_{01}/h_{02}=0.25`$ and $`h_{01}/h_{03}=0.1`$.
Fig.10. The same as in Fig. 8 for a varying relative detuning from the resonance $`\delta (\omega \omega _0)/\omega _0`$ taking the values: $`\delta =0`$ (solid line); $`\delta =0.025`$ (dashed line); and $`\delta =0.25`$ (solid line with squares).
Fig.11. Radiation intensivity $`I`$, coherence coefficient $`C_{coh}`$, and spin polarization $`p_z`$ versus time, in the absence of alternating external fields and with a weak coupling with a resonator, $`g1`$.
Fig.12. Phase portrait demonstrating a stable focus for the parameters $`z_0=0.5`$, $`w_0=0.001,g=10`$, and $`\gamma =0.1`$.
Fig.13. Pulsing regime of spin relaxation with the parameters $`z_0=0.1,w_0=10^6`$ and $`\gamma =0.01`$ for the functions: (a) $`w(t)`$; (b) $`z(t)`$.
Fig.14. The time dependence of the functions: (a) $`w(t)`$; (b) $`z(t)`$, for the parameters $`z_0=0.5,w_0=0.001`$, and $`\gamma =1`$.
Fig.15. Dynamics of slow solutions: (a) $`w(t)`$; (b) $`z(t)`$, for the parameters $`z_0=0.5,w_0=0.01`$, and $`\gamma =0.1`$.
Fig.16. Evolution of slow solutions: (a) $`w(t)`$; (b) $`z(t)`$, for the parameters $`z_0=0.5,w_0=0.01`$, and $`\gamma =0.01`$.
Fig.17. Temporal behaviour of the function $`w(t)`$ for different sets of parameters: (a) $`z_0=0.1,w_0=10^6,\gamma =0.001`$; (b) $`z_0=0.1,\omega _0=0.001,\gamma =0.01`$; (c) $`z_0=0.5,w_0=10^6,\gamma =0.1`$; (d) $`z_0=0.5,w_0=0.001,\gamma =0.01`$.
Fig.18. Function $`w(t)`$ for $`z_0=0.5,w_0=0.5,\gamma =1`$, and varying pump parameters: $`\zeta =0.5`$ (solid line); $`\zeta =0.3`$ (dashed line).
Fig.19. Electric current through the semiconductor surfaces in the case of $`a=0.1,Q_2=1,\gamma =1`$ and different mobilities: $`\mu _2=10`$ (solid line); $`\mu _2=5`$ (dashed line); $`\mu _2=3`$ (shortโdashed line). (a) Leftโsurface current $`J(0,t)`$; (b) Rightโsurface current $`J(1,t)`$.
Fig.20. Leftโsurface current $`J(0,t)`$ (solid line), rightโsurface current $`J(1,t)`$ (dashed line), and the total current $`J(t)`$ (shortโdashed line) for $`a=0.25,Q_2=0.1,\mu _2=10`$ and different relaxation parameters: (a) $`\gamma =1`$; (b) $`\gamma =10`$; (c) $`\gamma =25`$.
Fig.21. Total electric current $`J(t)`$ for $`a=0.25,Q_2=0.1,\mu _2=10`$ and varying relaxation parameters: $`\gamma =1`$ (solid line); $`\gamma =10`$ (dashed line); $`\gamma =25`$ (shortโdashed line).
Fig.22. Electric current through semiconductor for the parameters $`a=0.1,Q_2=1,\gamma =1`$ and different mobilities: $`\mu _2=10`$ (solid line); $`\mu _2=5`$ (dashed line); $`\mu _2=3`$ (shortโdashed line).
Fig.23. Electric current $`J(t)`$ as a function of time for $`a=0.1,\mu _2=3,\gamma =1`$ and different initial charges:: $`Q_2=0`$ (solid line); $`Q_2=0.25`$ (dashed line); $`Q_2=0.5`$ (shortโdashed line); $`Q_2=0.75`$ (dotted line); $`Q_2=1`$ (dashedโdotted line).
Fig.24. Electric current $`J(t)`$ for $`Q_2=1,\mu _2=3,\gamma =1`$, and different locations of initial charge layers: $`a=0.05`$ (solid line); $`a=0.1`$ (dashed line); $`a=0.15`$ (shortโdashed line); $`a=0.2`$ (dotted line); $`a=0.25`$ (dashedโdotted line).
Fig.25. Relative energy density as a function of temperature in MeV for the $`SU(3)`$ gluonโglueball mixture of different glueball radii: 0 (line 1); 0.5 fm (line 2); 0.7 fm (line 3); 0.8 fm (line 4); 1 fm (line 5).
Fig.26. Relative enthalpy for the gluonโglueball mixture as a function of temperature reduced to the deconfinement temperature, in the case of the glueball radius 0.82 fm, compared with the lattice numerical calculations.
Fig.27. Relative specific heat for the gluonโglueball mixture, for the glueball radius 0.82 fm, as a function of temperature in MeV.
Fig.28. Glueball channel probability versus temperature in MeV for the glueball radii as in Fig. 25.
Fig.29. Nucleon channel probability as a function of relative baryon density.
Fig.30. Dibaryon channel probability versus relative baryon density.
|
warning/0006/hep-ph0006234.html
|
ar5iv
|
text
|
# HIGH-ENERGY SCATTERING AND DIFFRACTION: THEORY SUMMARYaafootnote aTalk given at the 8th International Workshop on Deep-Inelastic Scattering (DIS2000), 25th-30th April 2000, Liverpool, England, to appear in the proceedings.
## 1 Introduction
On the one hand, small-coupling perturbation theory has been successfully applied to a variety of QCD processes. Its validity is well-understood in situations where intermediate states with high virtualities dominate. On the other hand, lattice Monte Carlo simulations provide a powerful first-principles approach to study the low-energy characteristics of the theory, such as the spectrum of hadronic excitations. However, there is still no established method, derived from the Lagrangian of QCD, that describes the high-energy scattering of hadrons. The reason for this is the difficulty to combine non-perturbative effects with the fundamentally Minkowskian physics in the high-energy limit. Thus, it can be argued that the high-energy limit represents one of the most interesting and difficult open problems in the theory of strong interactions. One obvious challenge is the derivation of the high-energy behaviour of hadronic cross sections, which are well-parametrised as $`\mathrm{ln}^2s`$ or $`s^{0.08}`$ (where $`\sqrt{s}`$ is the cms-energy of the collison), from the known microscopic theory.
Diffraction, and in particular the processes of hard diffraction discovered at the CERN S$`p\overline{p}`$S collider and studied in detail at HERA and the Tevatron, represent a powerful tool for the study of the high-energy limit of QCD. This is illustrated in Fig. 1, where forward Compton scattering, equivalent to the process of deep-inelastic scattering (DIS), is compared to diffractive electroproduction. Obviously, the study of different diffractive final states $`X`$ provides a wealth of hadronic high-energy scattering data, taking us far beyond the well-known inclusive process of DIS.
## 2 New Approaches to the High-Energy Limit of QCD
A fundamentally new approach to the high-energy limit of QCD has been advertised by Peschanski $`^{\mathrm{?},\mathrm{?}}`$. The authors suggest using the AdS/CFT correspondence (also known as the Maldacena conjecture) $`^\mathrm{?}`$ to investigate high-energy scattering in non-Abelian gauge theories. AdS/CFT correspondence claims the equivalence of weakly coupled string theory in an Anti-de-Sitter (AdS) geometry with strongly coupled $`๐ฉ=4`$ super Yang-Mills theory, which is a conformal field theory (CFT), in 4-dimensional Minkowski space. Further, to make the connection with the realistic case of confining gauge theories, the authors use Wittenโs proposal $`^\mathrm{?}`$ that a confining gauge theory is dual to string theory in an AdS black hole background. In the gauge theory, the high-energy scattering of two dipoles can be calculated from the correlation function of two Wilson loops. Using AdS/CFT correspondence, the calculation of the latter can be reduced to a minimal surface problem in an AdS black hole background. The results obtained so far show reggeization with unit intercept $`^\mathrm{?}`$.
A very different unconventional approach to high energy scattering has been suggested by Kharzeev and Levin $`^{\mathrm{?},\mathrm{?},\mathrm{?}}`$. They start from the leading order BFKL ladder diagram and emphasize the NLO contribution where the one-gluon rungs are replaced by pairs of gluons. Then, focussing on the soft region, these gluon pairs are replaced by pion pairs. The coupling to the vertical gluon lines of the ladder is fixed by employing the QCD anomaly relation
$$\theta _\mu ^\mu =\frac{\beta (g)}{2g}F^{a\alpha \beta }F_{\alpha \beta }^a$$
(1)
and calculating the trace of the energy momentum tensor $`\theta _\mu ^\mu `$ in terms of the pion degrees of freedom in chiral perturbation theory. After all ladder diagrams with two-pion rungs are summed, a soft-pomeron-like behaviour $`s^\mathrm{\Delta }`$ emerges. The intercept $`\mathrm{\Delta }=(1/48)\mathrm{ln}M_0^2/m_\pi ^2`$ comes out approximately right if the matching scale $`M_0^2`$, introduced by using chiral perturbation theory, is taken in the range $`4รท6`$ GeV<sup>2</sup>, in agreement with sum rule analyses.
As emphasized by Kharzeev, diffractive glueball production provides an interesting testing ground for this new picture of the pomeron $`^\mathrm{?}`$.
## 3 $`\gamma ^{}`$-$`\gamma ^{}`$ Scattering at High Energy
The total cross section of two highly virtual photons represents a unique testing ground for perturbative methods in high-energy scattering because the underlying process is the interaction of two small colour dipoles. The process is expected to include a kinematical region where the BFKL summation techiques are applicable.
Recent progress relevant to NLO BFKL calculations $`^\mathrm{?}`$ was discussed by Lipatov, who emphasized the enormous simplifications of the NLO BFKL kernel arising in $`๐ฉ=4`$ SUSY QCD and possible close relations between BFKL and DGLAP $`^{\mathrm{?},\mathrm{?}}`$. Furthermore, Lipatov noted the good description of $`\gamma ^{}\gamma ^{}`$ data in NLO BFKL achieved by using a non-Abelian physical renormalization scheme together with BLM scale fixing $`^\mathrm{?}`$. However, other methods to modify the naive NLO corrections to BFKL, which are extremely large, do also exist $`^\mathrm{?}`$.
The problems of the BFKL method justify the attempt to account for the data, which lies far above the Born term prediction, by other means. Naftali $`^{\mathrm{?},\mathrm{?}}`$ presented a calculation taking into account hard-soft and soft-soft contributions, which are also present in $`\gamma ^{}`$-$`\gamma ^{}`$ processes. Although significant enhancements were found, they are not sufficient to account for the data when both photon virtualities are large. Donnachie reviewed the recent phenomenological approach of the โtwo pomeronsโ $`^{\mathrm{?},\mathrm{?}}`$, which includes the well-known reggeon and soft pomeron trajectories and a phenomenological hard pomeron with an intercept $`0.44`$. This approach describes successfully $`\gamma `$-$`\gamma `$ and $`\gamma ^{}`$-$`\gamma `$ cross sections, but is below the data in the $`\gamma ^{}`$-$`\gamma ^{}`$ case.
## 4 Diffractive Electroproduction
A large part of the diffractive data at HERA can be characterized by the diffractive structure function $`F_2^D`$, which describes the process $`\gamma ^{}pp^{}X`$ (cf. Fig. 1). In addition to the conventional kinematic variables of DIS, $`Q^2`$ and $`x=x_{\text{Bj}}`$, the process is characterized by $`M`$, the mass of the diffractive final state $`X`$. Alternatively, the variables $`\beta =Q^2/(Q^2+M^2)`$ or $`\xi =x_{IP}=x/\beta `$ can be used. Now, $`F_2^{D(3)}(x,Q^2,\xi )`$ is defined precisely as $`F_2(x,Q^2)`$, but on the basis of a cross section that is differential in $`\xi `$ as well as in $`x`$ and $`Q^2`$. Elastic vector meson production, to be discussed in more detail below, is obtained by appropriately specifying the final state $`X`$. For recent theoretical reviews of diffractive DIS see refs. $`^{\mathrm{?},\mathrm{?}}`$.
Diffraction occurs if the hadronic fluctuation of the incoming virtual photon scatters off the proton without destroying its colour neutrality. At leading order, the fluctuation is a $`q\overline{q}`$ pair, and its interaction can be parametrized by the dipole cross section $`\sigma (\rho )`$, where $`\rho `$ is the transverse size of the dipole $`^\mathrm{?}`$. A QCD-improved parametrization of the dipole cross section which carefully implements its relation to the gluon distribution in the region of small $`\rho `$ and avoids unitarity violations associated with the strong growth of the gluon distribution at small $`x`$ was presented by McDermott $`^{\mathrm{?},\mathrm{?}}`$.
Diffractive parton distributions $`^\mathrm{?}`$, denoted here by $`df_i^D/d\xi `$, characterize the probability of finding a parton in the proton under the condition that the proton remains intact. In this framework, which is firmly rooted in perturbative QCD, the diffractive cross section reads
$$\frac{d\sigma (x,Q^2,\xi )^{\gamma ^{}pp^{}X}}{d\xi }=\underset{i}{}_x^\xi ๐y\widehat{\sigma }(x,Q^2,y)^{\gamma ^{}i}\left(\frac{df_i^D(y,\xi )}{d\xi }\right),$$
(2)
where $`\widehat{\sigma }(x,Q^2,y)^{\gamma ^{}i}`$ is the total cross section for the scattering of a virtual photon characterized by $`x`$ and $`Q^2`$ and a parton of type $`i`$ carrying a fraction $`y`$ of the proton momentum.
Royon presented a parametrization of $`F_2^D`$ as well as a QCD fit based on the DGLAP evolution of diffractive parton distributions $`^{\mathrm{?},\mathrm{?}}`$. A novel feature of this fit is the subtraction of higher twist contaminations at large $`\beta `$. It is interesting that the famous โpeakedโ gluon of previous H1 analyses $`^\mathrm{?}`$ seems to be disfavoured.
Schรคfer $`^\mathrm{?}`$ discussed results for $`F_2^{D(3)}`$ at small $`\beta `$, obtained in the colour-dipole Regge-expansion approach and stressed the relevance of unitarity corrections for the $`\xi `$ dependence.
Goulianos has suggested a simple parametrization $`^\mathrm{?}`$ of the $`F_2^D`$ data at HERA, which is based on the ansatz $`d^3\sigma /d\xi dxdQ^2F_2^h(x,Q^2)/x/\xi ^{1+ฯต}`$ with a โhardโ structure function $`F_2^h`$.
An important new result concerning the charm contribution to $`F_2^D`$ was presented by Bartels $`^\mathrm{?}`$. At leading order, diffractive charm production is realized by $`c\overline{c}`$ and $`c\overline{c}g`$ final states. Except for the large-$`\beta `$ region, the latter component dominates because it allows for soft colour-singlet exchange. The new results presented by Bartels extend previous calculations of $`c\overline{c}g`$ production, where the $`p_{}`$ of the gluon was assumed to be much smaller than the $`p_{}`$ of the quarks (strong $`p_{}`$ ordering), to general kinematic configurations excluding, however, the case of soft gluons.
## 5 Diffraction at Hadron-Hadron Colliders
A frequently discussed issue in hard diffractive processes where either one or both colliding hadrons remain intact is the question whether a simple connection with the partonic description of diffractive DIS can be found. As emphasized by Royon, the hadronic data undershoots HERA based expectations by a large factor $`^{\mathrm{?},\mathrm{?}}`$, which is indeed expected from the simple geometrical picture of the collision of two extended soft objects. Thus, a fundamentally new theoretical approach to this type of hadronic processes appears to be necessary. Timneanu $`^\mathrm{?}`$ reported the successful description of both HERA and Tevatron gap events by using a Monte Carlo implementation of a Soft-Colour-Interaction model based on the generalized area law. Also, as presented in the talk by Cox $`^{\mathrm{?},\mathrm{?}}`$, HERA and Tevatron data characterized by a gap between two jets can be described by LLA BFKL within the HERWIG Monte Carlo, if a fixed $`\alpha _S=0.17`$ is adopted and if multiple scattering for the underlying event is taken into account. The interesting process of Higgs (or dijet) production in double rapidity gap events was discussed by Khoze $`^\mathrm{?}`$. He presented refined calculations in a perturbative approach, which, however, lead to cross sections considerably smaller than those predicted by some non-perturbative models. Close explained $`^\mathrm{?}`$ how the pomeron can be studied in hadron collisions at low momentum transfer by measuring the $`\varphi `$ and $`t`$ dependence for different ($`J^{PC}=0^{\pm +},1^{++},2^{++}`$) mesons.
## 6 Elastic Meson Production
Elastic vector meson production $`\gamma ^{}pVp`$ is a rich field, both theoretically and exprimentally. For large photon virtualities $`Q^2`$ and/or heavy mesons (with large mass $`M_V`$) this process constitutes a nice laboratory to study diffractive hard scattering. At small $`\xi `$ the production amplitude factorizes in the fluctuation of the virtual photon, the elastic scattering of the $`q\overline{q}`$ (or $`q\overline{q}g,\mathrm{}`$ in higher orders) off the proton, and the formation of the final state (vector) meson ($`V`$): $`๐(\gamma ^{}pVp)=\psi _{q\overline{q}}^\gamma A_{q\overline{q}+p}\psi _{q\overline{q}}^V`$. At leading order the diffractive (colourless) exchange is realized by a pair of gluons $`^\mathrm{?}`$. In the usual collinear factorization approach the amplitude $`A_{q\overline{q}+p}`$ is then proportional to the gluon density in the proton $`\xi g(\xi ,\mu ^2)`$ at some effective scale $`\mu ^2(Q^2+M_V^2)/4`$.
Recent calculations as reported by I. Ivanov $`^\mathrm{?}`$ and Martin $`^\mathrm{?}`$ improve on these approximations (see also $`^\mathrm{?}`$). Firstly, the transverse momentum of the exchanged gluons is taken into account by applying the so-called $`k_T`$-factorization and using the unintegrated gluon distribution $`f(\xi ,k_T^2)`$, where $`k_T`$ is the transverse momentum of the exchanged gluons. Secondly, even in the case of forward scattering the need to transform a spacelike photon into a timelike vector meson forces the kinematics to be non-forward, and the usual parton (gluon) distributions have to be replaced by skewed (also called non-forward or off-diagonal) parton distribution functions (SPDF). These are generalizations of the normal PDFs (without direct probabilistic interpretation) and follow their own, new evolution equations. A method to construct corresponding exclusive evolution kernels at NLO was reported by Freund $`^\mathrm{?}`$. In this work explicit diagrammatic two-loop calculations are avoided by using conformal $`๐ฉ=1`$ SUSY Yang-Mills constraints together with known two-loop DGLAP kernels.
Within the perturbative two-gluon-picture, elastic (electro-) production of light and heavy vector mesons can be calculated in fair agreement with experimental data as long as $`Q^2`$ and/or $`M_V^2`$ provide a hard scale of several GeV<sup>2</sup>. As the cross section depends on the gluon distribution squared, the process $`\gamma ^{}pVp`$ may serve as a particularly sensitive probe of the gluon at small $`\xi `$. To achieve this ambitious goal the theoretical uncertainties should be decreased further. In addition use should be made of the different available observables, i.e., $`Q^2`$ and energy dependence of the total cross sections for different mesons, $`\sigma _L/\sigma _T`$, the ratio of longitudinal to transverse photon induced production, and maybe even the full spin density matrix of $`\rho `$ or $`J/\psi `$ production measurements. One particular source of uncertainty in the theoretical description is the meson wave function. As shown by I. Ivanov $`^\mathrm{?}`$ different wave function models lead to quite different results, especially for $`\sigma _T`$, and the wave function is expected to play a significant role in the production of excited vector mesons compared to ground states.
On the other hand, as demonstrated by Martin $`^\mathrm{?}`$, the basic features of elastic vector meson production are mainly controlled by the photon wave function and the gluon distribution and can therefore be well predicted in the framework of Parton-Hadron-Duality (PHD). This approach uses open $`q\overline{q}`$ pair production, integrated over an appropriate mass interval and projected on the quantum numbers of the meson under consideration, thus avoiding the meson wave function completely. Even $`\sigma _L/\sigma _T`$, which is poorly described in most other models, is in agreement with HERA data.
Another field where the use of perturbative QCD in the framework of the two gluon picture can be justified is diffractive meson production at large momentum transfer $`t`$. New interesting results were reported by D. Ivanov $`^{\mathrm{?},\mathrm{?}}`$. He obtained large contributions to high-$`t`$ light vector meson photoproduction from a โ normally highly suppressed โ chiral odd $`q\overline{q}`$ component, where the photon couples to the quarks via the magnetic susceptibility of the vacuum with a surprisingly large coefficient.
A particularly interesting possibility for observing the odderon, i.e., the $`C=P=1`$ partner of the pomeron, in the context of diffractive meson production has been suggested by Dosch $`^{\mathrm{?},\mathrm{?}}`$. Addressing the fundamental question why the odderon is not seen in the difference between $`pp`$ and $`p\overline{p}`$ cross sections, the authors suggest that the reason lies in the quark-diquark structure of the proton. If this is the case, then the diffractive production of pseudoscalar mesons at HERA with proton breakup in the final state should provide an ideal testing ground for the odderon. Employing a stochastic vacuum based approach in the description of the soft $`t`$-channel exchange $`^\mathrm{?}`$, a prediction of $`\sigma _{\gamma p\pi ^0X}300`$ nb with estimated model uncertainties of $`\pm 50\%`$ was given.
## 7 Conclusions and Outlook
The high-energy asymptotics of hadronic cross sections belong to the few phenomenologically relevant and not yet calculable implications of known quantum field theories. Thus, high-energy scattering and diffraction remain among the least well understood and therefore most interesting fields in QCD. Progress has been reported in may directions, with work ranging from simple phenomenology to elaborate multi-loop calculations. Given the fast and continuous improvement of data and the rich interplay between theory and experiment, we experience an exciting time for this fundamental area of research.
## Acknowledgements
We would like to thank the organizers for this very interesting and stimulating Workshop and the participants of our Working Group for their excellent presentations and for their enthusiasm during the discussions.
## References
|
warning/0006/hep-ph0006211.html
|
ar5iv
|
text
|
# Solving the Gravitino Problem by Axino
## Acknowledgements
This work was partially supported by the Japan Society for the Promotion of Science (T.A).
|
warning/0006/cond-mat0006017.html
|
ar5iv
|
text
|
# Atom loss and the formation of a molecular Bose-Einstein condensate by Feshbach resonance
## I Introduction
Most properties of a Bose-Einstein condensate (BEC) are determined by interatomic interactions (see Refs. ). These interactions are responsible for the characteristic nonlinear term in the equations of motion of the condensate, whose magnitude depends on the collisional elastic scattering length. Recent experiments on optically-trapped BECs drew attention to the effects of Feshbach resonances on the properties of the condensate, as scattering lengths are strongly modified by the presence of a resonance. More particularly, these experiments sought to modify these effects by application of a magnetic field (see Refs. ). One of the rather astonishing results observed was a dramatic loss of condensate population as the magnetic field was varied so that the ensuing Zeeman shift made the system pass through a resonance, or approach it closely.
A Feshbach resonance may exist when the energy of a pair of atoms in the condensate is close to that of a metastable molecular state Na$`{}_{2}{}^{}\left(m\right)`$. Then the scattering length varies strongly as a function of the energy mismatch between the two states. This mismatch can be controlled by applying a varying magnetic field. The energies of the two states can be brought closer to each other, as the two states have different Zeeman shifts. The MIT experiment strived to study the effect of the Zeeman shift by applying a time-varying magnetic field in two distinct procedures: (a) a fast sweep through the resonance, using a fast ramp speed of the magnetic field, and (b) a slow sweep, using much lower speeds, in which the ramp is stopped short of crossing the resonance. The latter procedure is then repeated by using different values of the stopping value of the magnetic field, and is carried out on both sides of the resonance. Both types of experiment resulted in a large condensate population loss.
This work is devoted to the study of possible mechanisms leading to this loss. Preliminary results were presented in Refs. , suggesting the mechanism of collisional deactivation. Another mechanism, involving trap excitation by a two-step curve-crossing process, was suggested in Refs. . We study below in more detail the combined effect of both mechanisms. As the MIT experiments were conducted with Na atoms, we shall refer here for definiteness to Na only, though this study may be extended to similar systems.
Given a stationary Zeeman shift, a population of molecules can be formed as a temporary stage in the elastic process
$`\text{Na}(\text{BEC})+\text{Na}(\text{BEC})\text{Na}_2\left(m\right)`$ (1)
$`\text{Na(BEC)}+\text{Na}(\text{BEC}).`$ (2)
However, in the absence of other intervening interactions, or of a time-varying field, this molecular population cannot persist, and no loss would occur.
The first loss mechanism considered here involves the deactivation of the resonance state, which is usually a highly excited vibrational level in a given spin state of the pair, by an exoergic collision with a third atom of the condensate ,
$$\text{Na}_2(m)+\text{Na}(\text{BEC})\text{Na}_2(d)+\text{Na}(\text{hot}),$$
(3)
bringing the molecule down to a stable state $`d`$, and releasing kinetic energy to the relative motion of the reaction products. Although the collision occurs with a vanishingly small kinetic energy, rates of such inelastic processes remain finite at near-zero energies . This process naturally depends on the initial density in the condensate. A variant of this process, involving deactivation by collisions with another molecule (rather than an atom), of the type
$$\text{Na}_2\left(m\right)+\text{Na}_2\left(m\right)\text{Na}_2\left(d\right)+\text{Na}_2\left(u\right).$$
(4)
would require a significant molecular density. The two molecules emerge in two states $`d`$ and $`u`$, where $`u`$ can be a stable molecular state above $`d`$, or a continuum state of a dissociating molecule. The reaction can take place as long as the corresponding internal energies obey the inequality $`E_d+E_u0`$, where the internal energy of $`m`$ serves as the zero reference point on the energy scale. A particularly effective reaction of type (4) would occur in the near-resonant case, in which $`0<E_u<|E_d|`$. A typical example, common in VV-relaxation, is that of $`v+v\left(v+1\right)+\left(v1\right)`$ (where $`v`$ is the vibrational quantum number of the state $`m`$). In this example, the kinetic energy is provided by the vibrational anharmonicity.
Both reactions (3) and (4) are thus exoergic, providing products with sufficient kinetic energy to escape the trap, as the characteristic transition energies exceed the trap depth. The kinetic energy may even be sufficient to produce an additional loss mechanism โ secondary collisions of the reaction products with condensate atoms (see Ref. ). The loss mechanism described here can be enhanced by bringing close to each other the energies of the two states involved in reaction (2) โ the BEC state of an atom pair and the resonant molecular state $`m`$ โ by the application of a time-varying Zeeman shift. It is not necessary that the energies of these states should cross. An actual crossing of the two states can cause an irreversible transfer of population from the condensate to the molecular states (see Refs. ). This crossing is also necessary, as the first step, in the other mechanism referred to earlier โ that of excitation of the trap states by a two-step crossing. The first crossing between the condensate and molecular states can occur in two directions, either by letting the molecular state move upwards in its energy, with respect to the condensate atom-pair state, or by letting it move downwards. In both cases the loss of the molecules can proceed via the deactivation mechanism. But only the upwards move alone can initiate the excitation mechanism. The second crossing, at a higher energy, then causes transfer of population to higher atomic states โ bound and continuum trap states in a condensate imbedded in an optical trap , or continuum states in a free condensate . This process is accompanied by an increase in energy
$$\text{Na}(\text{BEC})+\text{Na}(\text{BEC})\text{Na}_2\left(m\right)\text{Na}(\text{hot})+\text{Na}(\text{hot}).$$
(5)
(To be more precise, two-step excitation can in principle occur also in a downward move by the so-called โcounterintuitiveโ process in which the second crossing precedes the first one along the time axis in a $`Z`$-shaped formation . However, this effect is negligibly small in the present case.)
It should be made clear that both crossing directions have been taken into account in Ref. . However, that work does not specify what happens to the molecular population in the downward move. In principle, once the Zeeman shift undergoes the first crossing, the formation of a molecular population in the resonant state can be considered as a valid loss channel, no matter whether there are deactivating collisions or not. In that case, the loss rate should be independent of the deactivation rate. Our analysis here shows that this is generally not the case. In the case of the fast-sweep experiment, the loss would become independent of the deactivation rate only under certain conditions (referred to in Sec. II C 2 below as the โasymptoticโ conditions).
It is now understood that in the slow-sweep experiment the loss is produced almost exclusively by deactivating collisions, such as the atom-molecular collisions described by (3). We study here also the added effect of molecule-molecule collisions described by (4). In the case of the fast-sweep experiment, it has been claimed that the main cause of loss is due to the excitation processes of the kind described by (5), but it was shown that deactivating collisions cannot be discounted as a contributing mechanism.
This paper therefore aims to study the effect of combining the two kinds of mechanisms (condensate excitation vs. deactivating collisions) together. One of the major conclusions (discussed in Sec. IV below) is that the two processes may actually compete nonadditively with each other, rather then contribute additively to the loss process.
The paper begins, in Sec. II, by an expansion of the theoretical analysis used in to describe the effect of deactivating collisions. We show, among other things, how the coupled equations of the Gross-Pitaevskii type for the atomic and molecular condensates, introduced and studied earlier by Timmermans et al. , can be derived by an elimination of the product states of the deactivation process (the so-called โdumpโ states). The equations are then extended to include molecule-molecule collisions of type (4). In Sec. III we add to the deactivation model an effect representing the outcome of the excitation process. The results are presented and discussed in Sec. IV, in comparison with the MIT experiments.
## II The deactivation model
### A Hamiltonian and variational procedure
Let us consider an optically-trapped BEC exposed to an external homogeneous time-dependent magnetic field $`B\left(t\right)`$ used to tune a vibrationally excited molecular state $`m`$ to a Feshbach resonance with the state of a pair of unbound condensate atoms. In order to write down an hamiltonian for such a system, including the molecular dump states, we must introduce field annihilation operators of the atoms $`\widehat{\psi }\left(๐ซ\right)`$, of the molecular resonant state $`\widehat{\psi }_m^+\left(๐ซ_m\right)`$, and of the lower and upper dump states, $`\widehat{\psi }_d^+\left(๐ซ_m\right)`$ and $`\widehat{\psi }_u^+\left(๐ซ_m\right)`$, respectively \[see discussion following Eq. (4)\]. The hamiltonian can then be written as
$`\widehat{H}=`$ $`{\displaystyle d^3r\widehat{\psi }^+\left(๐ซ\right)\widehat{H}_a\widehat{\psi }\left(๐ซ\right)}+\widehat{V}_{el}`$ (8)
$`+{\displaystyle d^3r_m\underset{\alpha =m,u,d}{}\widehat{\psi }_\alpha ^+\left(๐ซ_m\right)\widehat{H}_\alpha \widehat{\psi }_\alpha \left(๐ซ_m\right)}`$
$`+\widehat{V}_h+\widehat{V}_h^++{\displaystyle \underset{d}{}}\left(\widehat{V}_d+\widehat{V}_d^+\right)+{\displaystyle \underset{ud}{}}\left(\widehat{V}_{ud}+\widehat{V}_{ud}^+\right).`$
The terms
$`\widehat{H}_a={\displaystyle \frac{1}{2m}}\widehat{๐ฉ}^2+V_a\left(๐ซ\right)\mu _aB\left(t\right),`$ (9)
(10)
$`\widehat{H}_\alpha ={\displaystyle \frac{1}{4m}}\widehat{๐ฉ}_m^2+V_\alpha \left(๐ซ_m\right)\mu _\alpha B\left(t\right),`$ (11)
(where $`\alpha =m,u`$, or $`d`$, includes the resonant state) are the hamiltonians for the noninteracting atoms and molecules. Here $`V_a\left(๐ซ\right)`$ and $`V_\alpha \left(๐ซ_m\right)`$ are the corresponding atomic and molecular energies as functions of the position $`๐ซ`$ of the atomic (or $`๐ซ_m`$ of the molecular) center of mass, whose values include the optical trap potentials and the position-independent differences of internal energies in the absence of the trap. Also, $`\mu _a`$and $`\mu _\alpha `$ are the corresponding magnetic moments.
Since the atoms and molecules are treated here as independent particles, the interaction responsible for the atom-molecule coupling \[reaction (2)\] can be written in the general form
$$\widehat{V}_h=d^3rd^3r^{}V_h\left(๐ซ๐ซ^{}\right)\widehat{\psi }_m^+\left(\frac{๐ซ+๐ซ^{}}{2}\right)\widehat{\psi }\left(๐ซ\right)\widehat{\psi }\left(๐ซ^{}\right),$$
(13)
in which the molecule, created as an independent particle preserves the position of the center of mass. However, considering that the interaction is localized within a range of atomic size, negligibly small compared to the condensate size and the relevant de Broglie wavelengths, one can use the approximation of zero-range interaction $`V_h\left(๐ซ๐ซ^{}\right)=g\delta \left(๐ซ๐ซ^{}\right)`$, and represent the interaction in the simpler form
$$\widehat{V}_h=gd^3r\widehat{\psi }_m^+\left(๐ซ\right)\widehat{\psi }\left(๐ซ\right)\widehat{\psi }\left(๐ซ\right).$$
(14)
The same arguments are applicable to the terms in Eq. (8) representing the deactivating collisions $`\widehat{V}_d`$ and $`\widehat{V}_{ud}`$ \[reactions (3) and (4), respectively\]. However, the use of a zero-range interaction would lead to a divergence in the ensuing calculations \[see discussion following Eq. (40) below\]. Therefore we keep these interactions as finite-range functions of the distance between the reaction products, writing
$`\widehat{V}_d={\displaystyle d^3rd^3r_m}`$ $`d_d\left(|๐ซ๐ซ_m|\right)\widehat{\psi }^+\left(๐ซ\right)\widehat{\psi }_d^+\left(๐ซ_m\right)`$ (16)
$`\times \widehat{\psi }\left({\displaystyle \frac{๐ซ+2๐ซ_m}{3}}\right)\widehat{\psi }_m\left({\displaystyle \frac{๐ซ+2๐ซ_m}{3}}\right)`$
$`\widehat{V}_{ud}={\displaystyle d^3r_1d^3r_2}`$ $`d_{ud}\left(|๐ซ_1๐ซ_2|\right)\widehat{\psi }_u^+\left(๐ซ_1\right)\widehat{\psi }_d^+\left(๐ซ_2\right)`$ (18)
$`\times \widehat{\psi }_m\left({\displaystyle \frac{๐ซ_1+๐ซ_2}{2}}\right)\widehat{\psi }_m\left({\displaystyle \frac{๐ซ_1+๐ซ_2}{2}}\right).`$
Here, as in (13) the position of the center of mass is preserved, but the finite-range nature of the interactions is retained in the functions $`d_d\left(\rho \right)`$ and $`d_{ud}\left(\rho \right)`$, whose actual shape will be discussed further down.
Finally, the part of the hamiltonian associated with elastic collisions (see Ref. ),
$`\widehat{V}_{el}={\displaystyle }d^3r[`$ $`{\displaystyle \frac{U_a}{2}}\widehat{\psi }^+\left(๐ซ\right)\widehat{\psi }^+\left(๐ซ\right)\widehat{\psi }\left(๐ซ\right)\widehat{\psi }\left(๐ซ\right)`$ (21)
$`+{\displaystyle \frac{U_m}{2}}{\displaystyle \underset{\alpha ,\alpha ^{}}{}}\widehat{\psi }_\alpha ^+\left(๐ซ\right)\widehat{\psi }_\alpha ^{}^+\left(๐ซ\right)\widehat{\psi }_\alpha ^{}\left(๐ซ\right)\widehat{\psi }_\alpha \left(๐ซ\right)`$
$`+U_{am}{\displaystyle \underset{\alpha }{}}\widehat{\psi }^+\left(๐ซ\right)\widehat{\psi }_\alpha ^+\left(๐ซ\right)\widehat{\psi }_\alpha \left(๐ซ\right)\widehat{\psi }\left(๐ซ\right)],`$
includes terms proportional to the zero-momentum atom-atom, molecule-molecule, and atom-molecule interactions,
$$U_a=\frac{4\pi \mathrm{}^2}{m}a_a,U_m=\frac{2\pi \mathrm{}^2}{m}a_m,U_{am}=\frac{3\pi \mathrm{}^2}{m}a_{am},$$
(22)
where $`a_a`$, $`a_m`$, and $`a_{am}`$ are the corresponding elastic scattering lengths, and the different numerical factors in the numerators reflect the different reduced masses.
We outline here the derivation of mean-field equations, involving $`c`$-number fields, that represent all the actively participating states listed above. This is accomplished by an extension of a well-known variational method (see Refs. ). Let us introduce the trial wavefunction
$`|\mathrm{\Phi }=`$ $`[1+{\displaystyle }d^3rd^3r_m{\displaystyle \underset{d}{}}\phi _d(๐ซ,๐ซ_m,t)\widehat{\psi }^+\left(๐ซ\right)\widehat{\psi }_d^+\left(๐ซ_m\right)`$ (24)
$`+{\displaystyle }d^3r_1d^3r_2{\displaystyle \underset{ud}{}}\phi _{ud}(๐ซ_1,๐ซ_2,t)\widehat{\psi }_u^+\left(๐ซ_1\right)\widehat{\psi }_d^+\left(๐ซ_2\right)]`$
$`\times `$ $`|\phi _0,\phi _m.`$ (25)
The factor
$`|\phi _0,\phi _m=\mathrm{exp}\{{\displaystyle }d^3r[`$ $`\phi _0(๐ซ,t)\widehat{\psi }^+\left(๐ซ\right)`$ (27)
$`+\phi _m(๐ซ,t)\widehat{\psi }_m^+\left(๐ซ\right)]\}|0`$
is a coherent state, formed by a product of exponential operators, involving atomic ($`\phi _0`$) and molecular ($`\phi _m`$) condensate states, and operating on the vacuum state $`|0`$. The linear factor preceding it in Eq. (25) includes the fields $`\phi _d(๐ซ,๐ซ_2,t)`$ and $`\phi _{ud}(๐ซ_1,๐ซ_2,t)`$ which are the correlated states of the products in reactions (3) and (4), respectively.
The linear form of the latter factor forces the trial wavefunction (25) to take into account only one-particle occupation of the non-resonant molecular states $`u`$ and $`d`$, as a constraint. This approximation is based on the assumption of fast removal of โhotโ particles from the trap. In contrast, the occupation of the resonant state ($`m`$) is allowed to reach the order of magnitude of the condensate-state occupation (as our calculations verify). Therefore $`\phi _m`$ describes a coherent molecular condensate.
Another constraint, that follows from the large energy difference between the dump states and the resonant state, is the condition
$$d^3r\phi _0^{}(๐ซ,t)\phi _d(๐ซ,๐ซ_m,t)=0$$
(28)
\[see discussion following Eq. (40) for justification\]. The trial wavefunction (25) can now be substituted into a variational functional (see Refs. )
$$\underset{\mathrm{}}{\overset{\mathrm{}}{}}๐t\frac{\mathrm{\Phi }|i\mathrm{}\frac{}{t}\widehat{H}|\mathrm{\Phi }}{\mathrm{\Phi }|\mathrm{\Phi }}.$$
(29)
Neglecting terms of the order of $`\phi _d^3`$ and $`\phi _u^3`$, and taking (28) into account, the use of a standard variational procedure (see Ref. ) then leads to a set of coupled equations for the atomic ($`\phi _0`$) and molecular ($`\phi _m`$) condensate fields (or โwavefunctionsโ), as well as for the dump states ($`\phi _d`$ and $`\phi _{ud}`$):
$`i\mathrm{}\dot{\phi }_0(๐ซ,t)=`$ $`\left(\widehat{H}_a+U_a|\phi _0(๐ซ,t)|^2+U_{am}|\phi _m(๐ซ,t)|^2\right)\phi _0(๐ซ,t)+2g^{}\phi _0^{}(๐ซ,t)\phi _m(๐ซ,t)+Q(๐ซ,t)\phi _m^{}(๐ซ,t)`$ (31)
$`i\mathrm{}\dot{\phi }_m(๐ซ,t)=`$ $`\left(\widehat{H}_m+U_{am}|\phi _0(๐ซ,t)|^2+U_m|\phi _m(๐ซ,t)|^2\right)\phi _m(๐ซ,t)+g\phi _0^2(๐ซ,t)`$ (33)
$`+Q(๐ซ,t)\phi _0^{}(๐ซ,t)+Q_m(๐ซ,t)\phi _m^{}(๐ซ,t)`$
$`i\mathrm{}\dot{\phi }_d(๐ซ_1,๐ซ_2,t)=`$ $`\left({\displaystyle \frac{1}{2m}}\widehat{p}_1^2+{\displaystyle \frac{1}{4m}}\widehat{p}_2^2E_d\right)\phi _d(๐ซ_1,๐ซ_2,t)+d_d\left(|๐ซ_1๐ซ_2|\right)\phi _0({\displaystyle \frac{๐ซ_1+2๐ซ_2}{3}},t)\phi _m({\displaystyle \frac{๐ซ_1+2๐ซ_2}{3}},t)`$ (34)
$`i\mathrm{}\dot{\phi }_{ud}(๐ซ_1,๐ซ_2,t)=`$ $`\left({\displaystyle \frac{1}{4m}}\widehat{p}_1^2+{\displaystyle \frac{1}{4m}}\widehat{p}_2^2E_uE_d\right)\phi _{ud}(๐ซ_1,๐ซ_2,t)+d_{ud}\left(|๐ซ_1๐ซ_2|\right)\phi _m^2({\displaystyle \frac{๐ซ_1+๐ซ_2}{2}},t),`$ (35)
where
$`Q(๐ซ,t)={\displaystyle d^3r_1d^3r_2\underset{d}{}d_d^{}\left(|๐ซ_1๐ซ_2|\right)\phi _d(๐ซ_1,๐ซ_2,t)\delta \left(๐ซ\frac{๐ซ_1+2๐ซ_2}{3}\right)}`$ (36)
$`Q_m(๐ซ,t)=2{\displaystyle d^3r_1d^3r_2\underset{ud}{}d_{ud}^{}\left(|๐ซ_1๐ซ_2|\right)\phi _m^{}(๐ซ,t)\phi _{ud}(๐ซ_1,๐ซ_2,t)\delta \left(๐ซ\frac{๐ซ_1+๐ซ_2}{2}\right)}.`$ (37)
The terms corresponding to elastic collisions, dependent on $`|\phi _d|^2`$ and $`|\phi _{ud}|^2`$, which should appear in Eqs. (31) and (33), are omitted since they are negligible compared to the terms including $`\phi _0`$ and $`\phi _m`$. In Eqs. (34) and (35) the terms corresponding to elastic collisions, Zeeman shifts, and trap potentials are neglected compared to the position-independent internal energies, denoted here $`E_d`$ and $`E_u`$.
### B Dump state elimination
The procedure used to eliminate the dump states is similar to the Weisskopf-Wigner method of the theory of spontaneous emission (see Ref. ). Equation (34) is of the form of a Schrรถdinger equation for two free particles with a source (the last term in the right-hand side). Such an equation can be solved by applying the Greenโs function method for free particles, with the result
$`\phi _d(๐ซ_1,๐ซ_2,t)=`$ $`{\displaystyle \frac{i}{\left(2\pi \right)^6\mathrm{}}}{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}๐t^{}{\displaystyle d^3Kd^3k\mathrm{exp}\left[i\left(\frac{\mathrm{}K^2}{6m}+\frac{3\mathrm{}k^2}{4m}\frac{E_d}{\mathrm{}}i0\right)\left(tt^{}\right)\right]}`$ (39)
$`\times \mathrm{exp}\left[i๐{\displaystyle \frac{๐ซ_1+2๐ซ_2}{3}}+i๐ค\left(๐ซ_1๐ซ_2\right)\right]{\displaystyle d^3\rho d_d\left(\rho \right)e^{i๐ค๐}d^3Re^{i\mathrm{๐๐}}\phi _m(๐,t^{})\phi _0(๐,t^{})}.`$
\[
Here $`๐`$ is the center-of-mass position of the three-atom system, $`๐`$ is the radius vector of the reaction products, and $`๐`$, $`๐ค`$ are the corresponding wavevectors. Since $`\phi _m(๐,t)`$ and $`\phi _0(๐,t)`$ are condensate wavefunctions, the Fourier transform of their product \[the integral over $`๐`$ in Eq. (39)\]vanishes if $`K>1/b`$, where $`b`$ is a characteristic size of the condensate. Therefore $`\mathrm{}^2K^2/\left(6m\right)<\mathrm{}\omega _{\text{trap}}`$ is negligible compared to $`E_d\mathrm{}\omega _{\text{trap}}`$, where $`\omega _{\text{trap}}`$ is the trap frequency. This fact allows us also to neglect the time dependence of $`\phi _m(๐,t)`$ and $`\phi _0(๐,t)`$ in the integration over $`t^{}`$, and thus obtain the simplified expression
$$\phi _d(๐ซ_1,๐ซ_2,t)=\frac{1}{\left(2\pi \right)^3}\phi _0(\frac{๐ซ_1+2๐ซ_2}{3},t)\phi _m(\frac{๐ซ_1+2๐ซ_2}{3},t)d^3kd^3\rho \frac{d_d\left(\rho \right)\mathrm{exp}\left[i๐ค\left(๐ซ_2๐ซ_1๐\right)\right]}{3\mathrm{}^2k^2/\left(4m\right)E_di0}.$$
(40)
\]
The atom-molecule pair is thus formed with a momentum $`\mathrm{}k_d=\sqrt{4mE_d/3}`$ of relative motion. The function $`\phi _d(๐ซ_1,๐ซ_2,t)`$ is a rapidly oscillating function of the coordinates, and therefore condition (28) is justified.
Substituting Eq. (40) into Eqs. (36), and (37) and introducing a Fourier transform of the function $`d_d\left(\rho \right)`$
$$\stackrel{~}{d}_d\left(k\right)=d^3\rho d_d\left(\rho \right)\mathrm{exp}\left(i๐ค๐\right),$$
(41)
one obtains
$$Q(๐ซ,t)=\left(\delta +i\mathrm{}\gamma \right)\phi _0(๐ซ,t)\phi _m(๐ซ,t),$$
(42)
where
$$\delta +i\mathrm{}\gamma =\frac{1}{\left(2\pi \right)^3}\underset{d}{}d^3k\frac{|\stackrel{~}{d}_d\left(k\right)|^2}{3\mathrm{}^2k^2/\left(4m\right)E_di0}.$$
(43)
Using the well-known identity $`\left(xi0\right)^1=๐ซx^1+i\pi \delta \left(x\right)`$, where $`๐ซ`$ denotes the Cauchy principal part of the integral, allows us to obtain explicit expressions for $`\gamma `$ and $`\delta `$:
$`\gamma `$ $`={\displaystyle \frac{m}{3\pi \mathrm{}^3}}{\displaystyle \underset{d}{}}k_d\left|\stackrel{~}{d}_d\left(k_d\right)\right|^2`$ (44)
$`\delta `$ $`={\displaystyle \frac{2m}{3\pi ^2\mathrm{}^2}}{\displaystyle \underset{d}{}}๐ซ{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐k{\displaystyle \frac{k^2|\stackrel{~}{d}_d\left(k\right)|^2}{k^2k_d^2}}.`$ (45)
A similar analysis, starting from Eq. (35), gives
$$Q_m(๐ซ,t)=\left(\delta _m+i\mathrm{}\gamma _m\right)\phi _m^2(๐ซ,t),$$
(46)
where
$`\gamma _m=`$ $`{\displaystyle \frac{m}{\pi \mathrm{}^3}}{\displaystyle \underset{u,d}{}}k_{ud}\left|\stackrel{~}{d}_{ud}\left(k_{ud}\right)\right|^2`$ (47)
$`\delta _m=`$ $`{\displaystyle \frac{2m}{\pi ^2\mathrm{}^2}}{\displaystyle \underset{u,d}{}}๐ซ{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐k{\displaystyle \frac{k^2|\stackrel{~}{d}_{ud}\left(k\right)|^2}{k^2k_{ud}^2}}.`$ (48)
and $`k_{ud}=\sqrt{2m\left(E_u+E_d\right)}/\mathrm{}`$. Substituting Eqs. (42) and (46) into Eqs. (31) and (33) one finally obtains a pair of coupled Gross-Pitaevskii equations (see Ref. )
$`i\mathrm{}\dot{\phi }_0`$ $`=\left(\widehat{H}_a+U_a|\phi _0|^2+U_{am}|\phi _m|^2\right)\phi _0+2g^{}\phi _0^{}\phi _m`$ (51)
$`\left(\delta +i\mathrm{}\gamma \right)|\phi _m|^2\phi _0`$
$`i\mathrm{}\dot{\phi }_m`$ $`=\left(\widehat{H}_m+U_{am}|\phi _0|^2+U_m|\phi _m|^2\right)\phi _m+g\phi _0^2`$ (53)
$`\left[\left(\delta +i\mathrm{}\gamma \right)|\phi _0|^2+\left(\delta _m+i\mathrm{}\gamma _m\right)|\phi _m|^2\right]\phi _m.`$
The parameters $`\delta `$, $`\gamma `$, $`\delta _m`$, and $`\gamma _m`$, which are expressed in terms of $`d_d`$ and $`d_{ud}`$ \[see Eqs. (44), (45), (47), and (48)\], describe the shift and the width of the resonance due to the deactivating collisions with atoms and molecules, respectively. The parameters $`\gamma `$ and $`\gamma _m`$ are one half of the corresponding rate constants. Since the strengths of the deactivating interactions are unknown, the parameter $`\gamma `$ will be extracted below from the experimental data, and $`\gamma _m`$ will be used below as an adjustable parameter. The shifts $`\delta `$ and $`\delta _m`$ can be incorporated in the interactions $`U_{am}`$ and $`U_m`$, respectively.
Equations (II B) are similar to those presented recently by Timmermans et al. . Among other things, Ref. shows that in the case of a time-independent magnetic field and large resonant detuning, neglecting the decay described by the imaginary terms, Eqs. (II B) can be reduced to a single Gross-Pitaevskii equation with an effective scattering length $`a_a\left[1\mathrm{\Delta }/\left(BB_0\right)\right]`$, where the parameter $`\mathrm{\Delta }`$ is related to the atom-molecule coupling constant $`g`$ of Eq. (14) as
$$|g|^2=2\pi \mathrm{}^2|a_a|\mu \mathrm{\Delta }/m.$$
(54)
Values of $`\mathrm{\Delta }`$ for Na were calculated in Refs. , or extracted from the experimental data in Refs. .
### C Density equations and approximate solutions
Let us introduce the new real variables
$`n(๐ซ,t)=|\phi _0(๐ซ,t)|^2,n_m(๐ซ,t)=|\phi _m(๐ซ,t)|^2,`$ (55)
$`u(๐ซ,t)=2\text{Re}\left(g\phi _0^2(๐ซ,t)\phi _m^{}(๐ซ,t)\right)/\mathrm{},`$ (56)
$`v(๐ซ,t)=2\text{Im}\left(g\phi _0^2(๐ซ,t)\phi _m^{}(๐ซ,t)\right)/\mathrm{}.`$ (57)
The time evolution of these variables can be described by a set of real equations, similar to the optical Bloch equations where $`n`$ and $`n_m`$ act like โpopulationsโ and $`u`$ and $`v`$ like โcoherencesโ. When the kinetic energy terms are neglected, in accordance with the Thomas-Fermi approximation (see Refs. ), one obtains from Eqs. (II B)
$`\dot{n}=2v2\mathrm{\Gamma }_an`$ (59)
$`\dot{n}_m=v2\mathrm{\Gamma }_mn_m`$ (60)
$`\dot{v}=Du\left(2\mathrm{\Gamma }_a+\mathrm{\Gamma }_m\right)v+2|g|^2n\left(4n_mn\right)/\mathrm{}^2`$ (61)
$`\dot{u}=Dv\left(2\mathrm{\Gamma }_a+\mathrm{\Gamma }_m\right)u.`$ (62)
Here
$`D(๐ซ,t)=\{V\left(๐ซ\right)\mu B\left(t\right)`$ (63)
$`+2\left[U_an(๐ซ,t)+\left(U_{am}\delta \right)n_m(๐ซ,t)\right]`$ (64)
$`[(U_{am}\delta )n(๐ซ,t)+(U_m\delta _m)n_m(๐ซ,t)]\}/\mathrm{},`$ (65)
$`V\left(๐ซ\right)=2V_a\left(๐ซ\right)V_m\left(๐ซ\right),\mu =2\mu _a\mu _m,`$ (66)
and
$`\mathrm{\Gamma }_a(๐ซ,t)=\gamma n_m(๐ซ,t),`$ (67)
(68)
$`\mathrm{\Gamma }_m(๐ซ,t)=\gamma n(๐ซ,t)+\gamma _mn_m(๐ซ,t).`$ (69)
In the Thomas-Fermi approximation the functions $`n(๐ซ,t)`$, $`n_m(๐ซ,t)`$, $`v(๐ซ,t)`$, and $`u(๐ซ,t)`$ depend on $`๐ซ`$ only parametrically. The set of four real equations (II C) can then be solved numerically using as initial conditions either an $`๐ซ`$-dependent (for example, a steady-state Thomas-Fermi) distribution, or an $`๐ซ`$-independent (homogeneous) distribution equal to the mean trap density.
Nevertheless, certain properties of the solutions can be derived analytically from Eqs. (II C), without recourse to numerical solutions, whenever the following โfast decayโ conditions hold:
$`\mu \dot{B}\mathrm{}\mathrm{\Gamma }_m\left(D+\mathrm{\Gamma }_m^2/D\right),\mathrm{\Gamma }_m\mathrm{\Gamma }_a`$ (71)
$`D^2+\mathrm{\Gamma }_m^26|g|^2n/\mathrm{}^2.`$ (72)
These conditions mean that the relaxation of $`n_m`$, $`v`$, and $`u`$ is much faster compared to that of $`n`$ and to the rate of change of the energy, caused by the magnetic field with a sweep rate $`\dot{B}`$. Therefore the values of the fast variables can be related to a given $`n`$ value, using a quasi-stationary approximation, by
$`u{\displaystyle \frac{D}{\mathrm{\Gamma }_m}}v,v{\displaystyle \frac{2|g|^2n^2\mathrm{\Gamma }_m}{\mathrm{}^2\left(D^2+\mathrm{\Gamma }_m^2\right)}},`$ (73)
(74)
$`n_m{\displaystyle \frac{|g|^2n^2}{\mathrm{}^2\left(D^2+\mathrm{\Gamma }_m^2\right)}},`$ (75)
and the condition (72) leads to $`n_mn`$. As a result, a single non-linear rate equation for the atomic density can be extracted. When terms proportional to the atomic and molecular densities in $`D`$ \[see Eq. (65)\] are neglected, the resulting rate equation is
$$\dot{n}(๐ซ,t)=\frac{6|g|^2\gamma n^3(๐ซ,t)}{[V\left(๐ซ\right)\mu B\left(t\right)]^2+\left[\mathrm{}\gamma n(๐ซ,t)\right]^2}.$$
(76)
(The neglected terms in $`D`$ effectively add an extra shift to the resonance, but its contribution is hardly noticed in the present problem.)
Equation (76) has a form analogous to the Breit-Wigner expression for resonant scattering in the limit of zero-momentum collisions (see Ref. ). In the Breit-Wigner sense one can interpret $`2\mathrm{}\gamma n`$ as the width of the decay channel, while the width of the input channel is proportional to $`|g|^2`$. This observation establishes a link between the macroscopic approach used here and microscopic approaches that treat the loss rate as a collision process. However, Eq. (76) differs from the usual Breit-Wigner expression by a factor $`\frac{3}{2}`$, associated with the loss of a third condensate atom in the reaction (3).
#### 1 Approaching the resonance
Very close to resonance (e.g., for conditions prevailing in the experiment , where $`B\left(t\right)`$ is within $`1\mu `$T of resonance) the behavior of Eq. (76) effectively attains a 1-body form linear in $`n`$. But as long as we stay out of this narrow region, by obeying the โoff-resonanceโ condition (which holds in the slow-sweep experiment ),
$$\mathrm{}\gamma n(๐ซ,t)|V\left(๐ซ\right)\mu B\left(t\right)|,$$
(77)
we can write Eq. (76) (to a very good approximation) in the 3-body form $`\dot{n}=K_3(๐ซ,t)n^3`$, where
$$K_3=\frac{12\pi \mathrm{}^2|a_a|\gamma \mathrm{\Delta }}{m\mu [B\left(t\right)V\left(๐ซ\right)/\mu ]^2}.$$
(78)
The dependence of Eq. (78) on the scattering length $`a_a`$ follows from Eq. (54). A similar expression has been obtained in Ref. , but the loss of the third condensate atom in the deactivating reaction (3), described by the term proportional to $`\mathrm{\Gamma }_a`$ in Eq. (59), was neglected. In the fast-decay approximation, in which $`\dot{n}_m`$ is diminishingly small, this neglected term is equal to $`v`$. Added to the $`2v`$, this makes $`K_3`$ of Eq. (78) larger by a factor $`\frac{3}{2}`$ than the corresponding expression in Ref. . This omission has been corrected in Refs. .
When the magnetic field ramp is assumed to vary linearly in time, starting at $`t_0`$ and ending at $`t`$, and Eq. (77) applies throughout the ramp motion (i.e., by avoiding passage through the resonance), the rate equation can be solved analytically. Using Eq. (78) one then gets
$`n(๐ซ,t)=n(๐ซ,t_0)[1+24\pi \mathrm{}^2|a_a|\mathrm{\Delta }\gamma n^2(๐ซ,t_0)`$ (79)
$`\times (tt_0)/\left(m\mu \dot{B}^2tt_0\right)]^{1/2},`$ (80)
where $`\dot{B}`$ is the magnetic-field ramp speed and the extrapolated time of reaching exact resonance is chosen to be $`t=0`$, so that both $`t`$ and $`t_0`$ have the same sign. We shall refer to the combination of Eqs. (II C) and (77), that leads to conditions (78), as the โthree-bodyโ approximation.
#### 2 Passing through the resonance
The three-body approximation of Eqs. (77) and (78) does not hold very close to resonance, and is therefore inapplicable to the description of the fast-sweep experiment, in which the Zeeman shift is swept rapidly through the resonance, causing dramatic losses (see Refs. ). Nevertheless, the fast decay approximation (II C) may still be valid. A simple analytical expression can then be derived for the condensate loss on passage though the resonance if, in addition, the magnetic field variation lasts long enough to reach the โasymptoticโ condition
$$\mu \delta B\mathrm{}\gamma n,$$
(81)
where $`\delta B`$ is the total change in $`B`$ accumulated over the sweep. This condition allows the extension of the ramp starting and stopping times to $`\mathrm{}`$.
The asymptotic behavior of $`n\left(t\right)`$ and $`\dot{n}\left(t\right)`$ in the complex $`t`$ plane, as $`|t|\mathrm{}`$, is constrained by consistency requirements. Consider the four possible asymptotic relations between $`n`$ and $`t`$ shown in the first column of Table I. Equation (76) forces $`\dot{n}`$ to attain the form in the second column, from which it follows that $`n`$ should attain the form in the third column. Obviously, cases (c) (for $`\text{Re}t>0`$) and (d) (at all $`t`$ values) are not self-consistent. Therefore, the asymptotic solution may attain only one of the forms complying with cases (a), (b), or (c) (the latter for $`\text{Re}t<0`$ only).
One can now evaluate the variation of $`n(๐ซ,t)`$ in the infinite time interval $`(\mathrm{},\mathrm{})`$. Let us rewrite Eq. (76) in the form
$$\frac{dn}{n^2}=\frac{6\gamma n|g|^2}{\left(\mu \dot{B}t\right)^2+\left(\mathrm{}\gamma n\right)^2}dt,$$
(82)
where $`V\left(๐ซ\right)`$ is removed by our choice of the origin on the time scale. We then integrate the left-hand side with respect to $`n`$ from $`n(๐ซ,\mathrm{})`$ to $`n(๐ซ,\mathrm{})`$ and the right-hand side with respect to $`t`$ from $`\mathrm{}`$ to $`\mathrm{}`$, considering $`n`$ as a well-defined function of $`t`$. The latter integral may be evaluated by using the residue theorem, closing the integration contour by an arc of infinite radius in the upper half-plane. The integral along this arc vanishes according to the asymptotic behavior of $`n`$ considered in all self-consistent cases of Table I.
The final result does not depend on $`\gamma `$, and on the self -consistent case studied, and has the form (valid for all positions $`๐ซ`$)
$`n(๐ซ,\mathrm{})={\displaystyle \frac{n(๐ซ,\mathrm{})}{1+sn(๐ซ,\mathrm{})}},`$ (83)
(84)
$`s={\displaystyle \frac{6\pi |g|^2}{\mathrm{}\mu |\dot{B}|}}={\displaystyle \frac{12\pi ^2\mathrm{}|a_a|}{m}}{\displaystyle \frac{\mathrm{\Delta }}{|\dot{B}|}}.`$ (85)
The product $`sn`$ in Eq. (84) would be proportional to the Landau-Zener exponent for the transition between the condensate and the resonant molecular states whose energies cross due to the time variation of the magnetic field, if one could keep the coupling strength $`g\phi _0`$ constant. However, for the non-linear curve crossing problem represented by Eq. (II B), the Landau-Zener formula is replaced by Eq. (84), which predicts a lower crossing probability, since the coupling strength $`g\phi _0`$ decreases, along with the decrease of the condensate density during the process.
The asymptotic result (84) describes the decay of the condensate density. Assuming a homogeneous initial density within the trap, Eq. (84) applies also to the loss of the total population $`N\left(t\right)=n(๐ซ,t)d^3r`$.
An asymptotic expression for the total population can also be found when the homogeneous distribution is replaced by the Thomas-Fermi one (see Ref. ). In this case, given $`n_0`$ is the maximum initial density in the center of the trap, one obtains
$`{\displaystyle \frac{N\left(\mathrm{}\right)}{N\left(\mathrm{}\right)}}={\displaystyle \frac{15}{2sn_0}}\{{\displaystyle \frac{1}{3}}+{\displaystyle \frac{1}{sn_0}}{\displaystyle \frac{1}{2sn_0}}\sqrt{1+{\displaystyle \frac{1}{sn_0}}}`$ (86)
$`\times \mathrm{ln}[(\sqrt{1+{\displaystyle \frac{1}{sn_0}}}+1)/(\sqrt{1+{\displaystyle \frac{1}{sn_0}}}1)]\}.`$ (87)
## III Inclusion of loss by excitation
The other mechanism of molecular condensate loss, considered in Ref. , involves the decay of the resonant molecular state by transferring atoms to excited (discrete) trap states or to higher-lying non-trapped (continuum) states. This decay is possible when the potential energy of the resonant state $`V_m\mu _mB\left(t\right)`$ exceeds the energy of the atom pair formed. A simpler version of this mechanism has been studied in Ref. .
Let us consider, for a moment, the decay of the molecule as a โhalf collisionโ, ignoring the finite size of the trap. The excess energy $`E`$ of the released pair, where $`E=E\left(t\right)=\mu B\left(t\right)V`$ at the time of release, determines a โhalf-widthโ of the Feshbach resonance. According to Eq. (27) of Ref. this energy-dependent half-width is given by
$$\mathrm{\Gamma }_{\text{cr}}\left(E\right)=\frac{|a_a|\mu \mathrm{\Delta }}{\mathrm{}^2}\sqrt{mE}.$$
(88)
It should be recalled that here always $`E>0`$. The coupling of the resonant molecular state with an excited state $`v`$ of an atom pair in the trap can be described by a coefficient $`g_v`$. This coefficient can be related to $`\mathrm{\Gamma }_{\text{cr}}`$ through (see Ref. )
$$g_v^2=\frac{\mathrm{}}{\pi }\mathrm{\Gamma }_{\text{cr}}\left(ฯต_v\right)\frac{ฯต_v}{v},$$
(89)
where $`ฯต_v`$ is the pair excited-state energy measured from the ground trap state and $`ฯต_v/v`$ measures the distance between the trap states in the vicinity of $`v`$ (i. e., the inverse density of states).
As the magnetic field $`B\left(t\right)=B_0+\dot{B}t`$ rises above the resonant value $`B_0`$, a crossing starts to occur between the resonant molecular state and the excited trap levels. Neglecting molecular collisions and motion, the resonant-state wavefunction $`\phi _m\left(t\right)`$ is propagated from $`\phi _m\left(t_0\right)`$ according to the rule (see Ref. )
$$\phi _m\left(t\right)=\phi _m\left(t_0\right)\mathrm{exp}\left(\frac{\pi }{\mathrm{}\mu \dot{B}}\underset{v}{}|g_v|^2\right).$$
(90)
Although $`g_v`$, for a given $`v`$ state, is time-independent, the sum taken over all $`v`$ states, bounded by the interval $`\mu \dot{B}t_0<ฯต_v<\mu \dot{B}t`$, is time-dependent. Whenever the amplitude of each crossing is small, i.e., when
$$|g_v|^2/\left(\mathrm{}\mu \dot{B}\right)1,$$
(91)
the sum in Eq. (90) can be replaced by an integral, giving
$$\phi _m\left(t\right)=\phi _m\left(t_0\right)\mathrm{exp}\left(\frac{\pi }{\mathrm{}\mu \dot{B}}\underset{t_0}{\overset{t}{}}|g_v|^2\frac{\mu \dot{B}}{ฯต_v/v}๐t\right),$$
(92)
where $`\left(ฯต_v/v\right)/\mu \dot{B}`$ measures the time interval between sequential crossings. Differentiation of Eq. (92) with respect to $`t`$ gives the following expression
$$\frac{\phi _m}{t}=\frac{\pi }{\mathrm{}}|g_v|^2\frac{v}{ฯต_v}\phi _m\mathrm{\Gamma }_{\text{cr}}\left(ฯต_v\right)\phi _m.$$
(93)
It should be kept in mind that this result is valid only in cases in which the energy gap between the atomic and molecular states increases rather fast; i.e., when $`\mu \dot{B}>0`$ and the condition (91) is obeyed. In the case of $`\mu \dot{B}<0`$, transitions from the ground trap state to excited ones are counterintuitive (see Ref. ), and become negligible when
$$\delta B\frac{g\sqrt{n}g_v}{\mathrm{}\mu \omega _{\text{trap}}},$$
(94)
where $`\delta B`$ is the range of variation of $`B`$ extended over both sides of the resonance.
Thus, whenever the energy gap between the resonant molecular and condensate states is positive, it increases rather fast \[following Eq. (91)\], and Eq. (94) is obeyed, one can account for the decay of the resonant molecular state into excited trap states by adding a term $`i\mathrm{}\mathrm{\Gamma }_{\text{cr}}\left(\mu \dot{B}t\right)\phi _m`$ to the right hand side of Eq. (53) or by substituting $`\mathrm{\Gamma }_m=\gamma n+\gamma _mn_m+\mathrm{\Gamma }_{\text{cr}}\left(\mu \dot{B}t\right)`$ in Eqs. (II C). In this way, the two mechanisms can be combined in a single formalism. Calculations made using this formalism are discussed in the following section.
## IV Results and Discussion
Calculations have been carried out on the loss of atoms for the case of the Na BEC experiments using both analytical results \[Eqs. (78), (80), (84), and (87)\] and numerical solutions of Eq. (II C). The parameters used to describe the system, reported previously , are $`a_a`$ = 3.4nm, $`\mu =2\mu _a\mu _m=3.65\mu _B`$ (where $`\mu _B=9.27\times 10^{24}`$J/T is the Bohr magneton), and $`\mathrm{\Delta }=`$ 0.95$`\mu `$T and 98$`\mu `$T, respectively, for the two resonances observed at 85.3mT (853G) and 90.7mT (907G) . These values of $`\mathrm{\Delta }`$ agree with the measured value for the 90.7mT resonance , and with the indirectly inferred order of magnitude for the 85.3mT resonance . The scattering lengths $`a_m`$ and $`a_{am}`$ for molecule-molecule and atom-molecule collisions, respectively, are not known, but calculations show that the results of our analysis are practically insensitive to the variation of their values as long as they stay within the order of magnitude of $`a_a`$.
In the calculations one should make a clear distinction between the two types of experiments conducted at MIT โ the slow-sweep and the fast-sweep experiments. In the first case, the values of the magnetic field at which the ramp stops short of resonance are such that the conditions needed for the excitation mechanism to occur do not exist, and only the collisional deactivation applies.
The graphs shown in Figs. 7 and 7 pertain to the slow-sweep MIT experiment for the strong 90.7 mT resonance. This resonance has been approached from below with two ramp speeds, and from above with one. Figure 7 shows the surviving atomic density $`n`$, and Fig. 7 shows $`K_3`$, both vs. the stopping value of the magnetic field $`B`$. The difference between Eqs. (78), (80) and the results of a direct numerical solution of Eqs. (II C) for all ramp speeds is so small that the corresponding plots are indistinguishable. The calculated plots were obtained using homogeneous-density initial conditions, starting from a $`B`$ value of 89.4mT on approach from below and 91.6mT from above. The corresponding initial mean densities were extracted from the experimental data . A best fit of the parameter $`\gamma `$, using Eq. (80), to the MIT data gives $`\gamma =0.8\times 10^{10}`$ cm$`{}_{}{}^{3}/`$s (see Fig. 7). But owing to the large scattering of the experimental data, and the associated uncertainty in the value of $`\gamma `$, we proceeded using the value of $`10^{10}`$cm$`{}_{}{}^{3}/`$s in our calculations, following Ref. . Given a density of about $`10^{15}`$cm<sup>-3</sup> this value of $`\gamma `$ implies a deactivation time of $`10^5`$s. The molecule-molecule deactivating collisions are negligible compared to the atom-molecule collisions due to the small molecular density \[see Eq. (74)\] whenever the fast decay condition (II C) holds.
The rate of condensate loss due to atom-molecule deactivating collisions was also studied in Ref. . As noted in the previous section, the expression for $`K_3`$ obtained there is smaller by a factor 1.5 from that of Eq. (78) \[see discussion after this equation\]. Without this factor one would not obtain the almost perfect match between the analytical and the numerical results mentioned in the discussion of Figs. 7 and 7 above. This omission has been corrected in Ref. and the value they obtain for their $`G_{\text{stab}}`$(corresponding to our $`2\gamma `$), of $`4\times 10^{10}`$ cm$`{}_{}{}^{3}/`$s, was estimated by best agreement with the experimental data . This value is 2.5 times bigger than our estimate. This discrepancy may be due to the large scatter in the experimental data. The estimate of Ref. is based only on the experimental $`K_3`$ plot (see Fig. 7) which is obtained by a differentiation of the experimental density data, and therefore shows a scatter of the data of up to an order of magnitude (see Fig. 7, as well as the corresponding figure in Ref. , that use a logarithmic scale). Equation (80) (presented in Ref. ) allows us to estimate the value of $`\gamma `$ directly from the experimental plot for the atomic density (see Fig. 7), that shows a much smaller ($`20\%`$) scatter of the data points. Anyhow, in both cases, the error bar should be comparable to the value of $`\gamma `$ itself.
An inelastic rate coefficient $`2\gamma `$ with a magnitude of the order of $`10^{10}`$cm$`{}_{}{}^{3}/`$s appears to be reasonable. First, this value is two orders of magnitude smaller than the upper bound set by the unitarity constraint on the $`S`$-matrix . In the limit of small momentum, unitarity provides $`2\gamma \mathrm{}\lambda /m`$, where $`\lambda `$ (the de Broglie wavelength) in the current situation is limited by the experimental trap dimensions. This constraint sets an upper bound of $`2.5\times 10^8`$cm$`{}_{}{}^{3}/`$s to $`2\gamma `$. Second, our estimate of $`10^{10}`$cm$`{}_{}{}^{3}/`$s for $`\gamma `$ is consistent with the order of magnitude of recently calculated vibrational deactivation rate coefficients due to ultracold collisions of He with H<sub>2</sub> in highly excited vibrational levels.
The remaining figures (7 to 7) pertain to the fast-sweep MIT experiment. Figures 7 and 7 present the surviving part of the trap population after passing through each of the two resonances. Following the experimental conditions, the value of $`10^{15}`$ cm<sup>-3</sup> is used for the initial density, and the magnetic field starting and stopping values are shifted from resonance by 3 mT for the strong (90.7 mT) resonance and by 2 mT for the weak (85.3 mT) one.
The analytical results of Eq. (84), together with the direct numerical solutions of Eqs. (II C) for the homogeneous initial distribution, considering only the deactivating mechanism, are compared in Fig. 7 with the results of the fast-sweep experiment . Although Fig. 7 does not specify the direction of the Zeeman shift $`\mu \dot{B}`$, one should recall (see Secs. I and II C 2 above) that the excitation mechanism can be ignored when $`\mu \dot{B}<0`$. The asymptotic result (84) reproduces a characteristic dependence on the ramp speed, although it is independent of $`\gamma `$. The numerical solutions clearly show a dependence on $`\gamma `$, as assumptions (II C) and (81) underlying the asymptotic result (84) do not hold in this case. The loss reaches a maximum at a value of $`\gamma `$ dependent on the various parameters (e.g., $`\gamma 10^{10}\text{cm}^3/\text{s}`$ for the 90.7 mT resonance when $`sn_02`$), as a result of the conflicting asymptotic and fast-decay conditions (II C) and (81). The calculated drop in the condensate loss on increase of the molecular dumping rate $`\gamma _m`$, which may seem paradoxical, can have the following explanation. Reaction (3) leads to the loss of three condensate atoms per each resonant molecule formed, while reaction (4) leads to the loss of only two condensate atoms. In the present case, the loss rate is limited by resonant molecule formation \[reaction (2)\]. Therefore, the increase of $`\gamma _m`$ transfers flow from channel (3) to channel (4), thus reducing the number of condensate atoms lost per each resonant molecule formed.
The results of the calculations, in which the effect of crossing to excited trap states is incorporated, are plotted in Fig. 7. The crossing rate was calculated with Eq. (88), using the values of $`\mathrm{\Delta }`$, $`a_a`$, and $`\mu `$ given at the beginning of this Section. The results clearly show that the two mechanisms do not augment each other. Adding the two-body excitation mechanism to the more efficient three-body deactivation mechanism actually reduces the loss. A possible explanation of this paradoxical result is similar to the one used in explaining Fig. 7.
Our predicted losses are somewhat higher than the ones obtained in the experiments for the 85.3 mT resonance, but significantly lower for the 90.7 mT resonance. Actually, the closest we can get to the experimental results is by considering only the two-body excitation mechanism for the 85.3 mT resonance, and the three-body deactivation mechanism for the 90.7 mT resonance.
The results of Ref. , considering only the two-body mechanism, also show a better agreement for the 85.3 mT resonance. There are, however, significant differences between the theory used there and the one presented here and in Ref. . In Ref. the parameter $`\gamma _0`$ (analogous to our $`\mathrm{\Gamma }_{\text{cr}}`$) is considered as a constant, independent of the released energy $`E`$ or time $`t`$, and prevailing all along the sweep, below and above the resonance. The time integral in their Eq. (4) should be smaller by a factor of 2 if the correct energy dependence of $`\gamma _0`$ is used. In our version, following Ref. , $`\mathrm{\Gamma }_{\text{cr}}`$ is $`t`$ -dependent, exists only for $`E>0`$, and attains the correct Wigner limit when $`E0`$. The final result of Ref. is an expression similar to our Eq. (84), with the exception that our parameter $`s`$ is larger by a factor $`3/2`$ than the corresponding parameter in Ref. . This difference reflects the fact that Eq. (84) was derived for the three-body deactivation process, while Ref. deals exclusively with a two-body excitation process. As a matter of fact, their coefficient should be further reduced by a factor of 2 associated with the energy dependence discussed above. Therefore the mechanism of Ref. requires a factor $`s`$ 3 times smaller than the one used in Eq. (84) for the deactivation mechanism.
Our results also differ from those of Ref. . The present calculations take into account the decrease of the condensate density during the crossing \[see discussion after Eq. (84)\] and therefore produce a lower condensate loss than the one obtained in Ref. , in which the loss is described by a Landau-Zener formula. In the limit of a small loss, Eq. (84) may resemble a Landau-Zener expression in which $`sn`$ is substituted for the exponent. However, a study of the associated free-atom problem shows that, even when we retain the initial density $`n(๐ซ,\mathrm{})`$ in the exponent, the latter would still be smaller than $`sn`$ by a factor of 3. This factor is directly related to the many-body character of the Gross-Pitaevskii equations.
At the first stage of the atomic condensate loss (2) a condensate of molecules in the resonant state is formed. This molecular condensate is unstable, due to the deactivation of the molecules by reactions (3) and (4), as well as their decay through the excitation mechanism. Figure 7 presents the calculated time dependence of the molecular condensate density, for various magnetic field ramp speeds, taking into account the various loss mechanisms. The oscillations of the molecular condensate density are connected to the intercondensate tunneling considered in Ref. . Figure 7 shows that the excitation loss enhances the damping of the oscillations and the decay of the molecular density. Nevertheless, the molecular condensate persists at least a few tenths of a microsecond before decaying. This time is long enough to allow converting the population of the vibrationally-excited state $`m`$ to the ground molecular state by methods of coherent control . Various techniques exist today for transfering populations coherently to a preselected state . The choice of technique would be dictated by properties of the process and of the target state (such as selection rules and stability).
The calculated peak values of the resonant molecular state density are presented in Figs. 7 and 7 for various rates of the deactivating collisions. Figure 7 refers only to the deactivation mechanism involving atom-molecule and molecule-molecule collisions, which take place for $`\mu \dot{B}<0`$, whereas Fig. 7 takes into consideration the combined effect of the two mechanisms (deactivation and excitation) which may take place for $`\mu \dot{B}<0`$. These figures show that from about 10% to 90% of the atomic density can be converted temporarily to a molecular condensate in the resonant state, in spite of losses due to the excitation mechanism and the molecule-molecule deactivation collisions. The calculated molecular condensate density is higher for the 90.7 mT resonance, or when the magnetic field ramp speed is not too large.
## V Conclusions
This paper discusses the two types of loss experiments conducted at MIT on sodium BEC, using a time-varying magnetic field in the proximity of a Feshbach resonance. The various processes discussed here involve a temporary formation of a molecular condensate. In the slow-sweep experiment the dominant loss mechanism is three-body deactivation by atom-molecule inelastic collisions. A best fit of an analytical expression for the atomic density obtained here \[Eq. (80)\] to the MIT data yields a rate coefficient for deactivating atom-molecule collisions of $`2\gamma =1.6\times 10^{10}`$ cm$`{}_{}{}^{3}/`$s. For the fast-sweep experiment an analytical expression, generalizing the Landau-Zener formula to a case of coupled nonlinear equations, is obtained. A different two-body mechanism, involving an excitation of the condensate by curve crossing, has been proposed previously . The combined effect of the two mechanisms is studied here, including also molecule-molecule deactivating collisions. The analysis shows that both processes should be taken into account, that they do not contribute additively to the loss, and that the outcome of the competition between them varies from one Feshbach resonance to another. Our numerical results show that, under favorable conditions, a substantial fraction of the trap population is converted to an unstable molecular condensate. This condensate persists long enough to allow its coherent transfer to a more stable state.
|
warning/0006/quant-ph0006098.html
|
ar5iv
|
text
|
# Unconditional Pointer States from Conditional Master Equations
## Abstract
When part of the environment responsible for decoherence is used to extract information about the decohering system, the preferred pointer states remain unchanged. This conclusion โ reached for a specific class of models โ is investigated in a general setting of conditional master equations using suitable generalizations of predictability sieve. We also find indications that the einselected states are easiest to infer from the measurements carried out on the environment.
PACS numbers: 03.65.Bz, 03.65.-w, 42.50.Lc
Introduction. โ Open quantum systems undergo environment-induced superselection (einselection) which leads to a preferred set of quasi-classical pointer states . They entangle least with the environment โ and therefore, lose least information. Hence, they can be found using predictability sieve, which seeks states minimizing entropy production .
However, the information lost to the environment could be, in principle, intercepted and recovered. Will the preferred states remain at least approximately the same when the environment is monitored in this fashion? This is a serious concern, as decoherence is caused by the entanglement between the system $`๐ฎ`$ and the environment $``$. It is well known that a pair of entangled quantum systems suffers from the basis ambiguity: One can find out about different states of one of them (e. g., $`๐ฎ`$) by choosing a different measurement of the other (e. g., $``$ ) . Thus, basis ambiguity may endanger the definiteness of the einselected states.
This issue was pointed out, for example, by Carmichael et al. , who used complete monitoring of the photon environment to develop a trajectory approach to quantum dynamics . Ref. demonstrated that โ when all of $``$ can be intercepted โ any basis of $`๐ฎ`$ can be inferred from the appropriate measurement on $``$, so at least in that limit substantial ambiguity is inevitable. This concern is further underscored by the realization that nearly all of our information comes not from direct observation of the system, but, rather, by intercepting a small fraction of (e. g. photon) environment.
Here we use predictability sieve in combination with the conditional master equation (CME) (which obtains when only a part of the environment โ and not all of it โ is traced out). We show โ using specific models โ that even when the additional data are taken into account, the pointer states are unchanged. We demonstrate, using fidelity, that even when all of $``$ is intercepted pointer states are unchanged. Moreover, using specific models we find indications that โ for an observer who acquires the data about the system indirectly by monitoring the environment โ pointer states are easiest to discover.
An example of CME.โ The master equation for a driven two-level atom whose emitted radiation is measured by homodyne detection is an example of CME,
$$d\rho =d\rho ^{\mathrm{UME}}[\rho ]+d\rho _{\mathrm{st}}[\rho ,N];$$
(1)
$`d\rho ^{\mathrm{UME}}=idt[\mathrm{\Omega }\sigma _x,\rho ]+dt\left(c\rho c^{}{\displaystyle \frac{1}{2}}c^{}c\rho {\displaystyle \frac{1}{2}}\rho c^{}c\right),`$ (2)
$`d\rho _{\mathrm{st}}=(dN\overline{dN})\left({\displaystyle \frac{(c+\gamma )\rho (c^{}+\gamma ^{})}{\mathrm{Tr}[(c+\gamma )\rho (c^{}+\gamma ^{})]}}\rho \right),`$ (3)
where we set the spontaneous emission time to $`1`$. We use the Itรด version of stochastic calculus. $`\rho `$ is a $`2\times 2`$ density matrix of the atom, $`\mathrm{\Omega }`$ is a frequency of transitions between the excited and the ground state driven by a laser beam, $`\gamma =Re^{i\varphi }`$ is the amplitude of the local oscillator in the homodyne detector and $`c=(\sigma _xi\sigma _y)/2`$ is an annihilation operator. $`N_t`$ is the number of photons detected until time $`t`$. Its increment $`dN\{0,1\}`$ is a dichotomic stochastic process with the average
$`\overline{dN}`$ $`=`$ $`\eta dt\mathrm{Tr}[\rho (c^{}+\gamma ^{})(c+\gamma )]`$ (4)
$`=`$ $`\eta dt[R^2+\sigma _xR\mathrm{cos}\varphi \sigma _yR\mathrm{sin}\varphi +c^{}c]`$ (5)
and $`\overline{dN^2}=\overline{dN}`$. The parameter $`\eta [0,1]`$ describes the efficiency of the measurement. Fully efficient measurement ($`\eta =1`$) occurs when the observer is continuously projecting the environment onto a pure state, so that an initial pure state of the system remains pure after the measurement . In the fully inefficient $`\eta =0`$ case, the observer has no measurement records or ignores them completely. In this case, $`d\rho _{\mathrm{st}}=0`$. The equation reduces to the unconditional master equation (UME), $`d\rho =d\rho ^{\mathrm{UME}}`$.
The average over realisations $`\overline{d\rho }=d\rho ^{\mathrm{UME}}`$, because $`\overline{dN\overline{dN}}=0`$ and $`\overline{d\rho _{\mathrm{st}}}=0`$. $`\overline{d\rho }=d\rho ^{\mathrm{UME}}`$ is not a special property of Eqs.(1,3) but an axiomatic property of any CME. The noise average means that we ignore any knowledge about the state of environment so the state of the system cannot be conditioned by this knowledge.
For $`R=0`$ the measurement scheme is simply a photodetection: $`\overline{dN}`$ is proportional to the probability that the atom is in the excited state. Every click of the photodetector ($`dN=1`$) brings the atom to the ground state, from where it is excited again by the laser beam. For $`R1`$ the homodyne photodetector current is a linear function of $`\sigma _x\mathrm{Tr}(\rho \sigma _x)`$ for $`\varphi =0`$ ($`x`$measurement) or of $`\sigma _y`$ for $`\varphi =\pi /2`$ ($`y`$measurement). These homodyne measurements drive the conditional state of the atom towards $`\sigma _x`$ and $`\sigma _y`$ eigenstates respectively.
Conditional pointer states are unconditional. โ Are the pointer states of a stochastic CME the same as pointer states of its corresponding deterministic UME ? An affirmative answers requires the assumption that there is only one kind of environment coupled to the system and to the detector, but to nothing else, so that the detector can, in principle, be fully efficient ($`\eta =1`$) in continuously projecting the environment onto pure states. This assumption is standard in quantum optics .
According to predictability sieve , pointer states minimize the increase of von Neumann entropy, or, equivalently, the decrease of purity $`P=\mathrm{Tr}(\rho ^2)`$ due to the interaction with an environment. Suppose that we prepare a system in a pure state $`\rho _0=\rho _0^2`$. The noise-averaged initial rate of purity loss is
$`\overline{dP}_0\mathrm{Tr}(2\rho _0\overline{d\rho _0})+\mathrm{Tr}(\overline{d\rho _0d\rho _0}).`$ (6)
Any CME can be written in the form of Eq.(1), where $`N_t`$ would represent a general stochastic process. The stochastic process feeds the information from measurements of $``$ into the conditional state of $`๐ฎ`$. The noise-averaged $`\overline{d\rho _0}=d\rho _0^{\mathrm{UME}}`$ depends neither on the efficiency $`\eta `$ nor on the kind of measurement we make on $``$. For a deterministic UME the second term on the RHS of Eq.(6) would be $`O(dt^2)`$. For a stochastic CME this second term gives a contribution proportional to $`\eta dt`$ which comes from $`\mathrm{Tr}[\overline{d\rho _{\mathrm{st}}d\rho _{\mathrm{st}}}]`$. The manifestly positive second term reduces the rate of purity loss because a measurement of $``$ tends to purify the conditional state. For $`\eta =1`$ the observer gains full knowledge about the environmental state, the conditional state of the system remains pure all the time, and $`dP_t=0`$. Thus we see that for $`\eta =1`$ the two terms of Eq.(6) should cancel one another. Given that the first and the second term cancel for $`\eta =1`$ and that the second term is linear in $`\eta `$, we can write the initial purity loss rate as
$$\overline{dP}_0=(1\eta )\mathrm{Tr}(2\rho _0d\rho _0^{\mathrm{UME}}).$$
(7)
Up to the prefactor of $`(1\eta )`$ this expression is the same as the corresponding one for the UME. Except for $`\eta =1`$ we can conclude that pointer states are the same as those for the UME no matter what the efficiency is or what kind of measurement is being made.
When $`\eta =1`$ we have $`\overline{dP}_0=0`$ and no preferred pointer states can be distinguished with the predictability sieve, in accordance with . However, even the conditional pure state can drift away from the free unitary evolution due to the coupling with $``$ which is measured completely ($`\eta =1`$). The faster it drifts away the less predictable the state of the system is. The fidelity with respect to the initial state is defined as $`F_t=\mathrm{Tr}(\rho _0\rho _t^{\mathrm{int}})`$, where the superscript int refers to interaction picture. For any $`\eta `$ the noise-averaged initial decrease of fidelity is
$$\overline{dF}_0\mathrm{Tr}(\rho _0\overline{d\rho _0})=\mathrm{Tr}(\rho _0d\rho _0^{\mathrm{UME}}).$$
(8)
Thus, UME pointer states maximize fidelity.
Fidelity and purity provide a basis for two physically different criteria which lead to the same unconditional pointer states . With the benefit of the hindsight, these results are not totally unexpected: UME is an average over CMEโs, so for linear predictability criteria pointer states should not change. We have however seen that the same holds for purity, which is non linear.
The expressions (7,8) can be worked out for the example of the two-level atom master equation. For $`\mathrm{\Omega }1`$ there is one pointer state: the ground state. An atom in the ground state cannot change its state by photoemission and the external driving is slow. In the limit $`\mathrm{\Omega }1`$ the externally driven oscillations are much faster than photoemission. In fact it would be misleading to use Eqs. (7,8) as they stand. It is more accurate to average them over one period of oscillation: $`\overline{dF}_0\frac{\mathrm{\Omega }}{2\pi }_0^{2\pi /\mathrm{\Omega }}๐t\mathrm{Tr}[\rho _0c_{\mathrm{int}}^{}\rho _0c_{\mathrm{int}}\rho _0c_{\mathrm{int}}^{}c_{\mathrm{int}}]=(3+x_0^2)/8`$. Here the density matrix is parametrized by $`\rho _t=[I+x_t\sigma _x+y_t\sigma _y+z_t\sigma _z]/2`$ with $`x_t^2+y_t^2+z_t^21`$. Given the last constraint, the states with $`x=\pm 1`$ (eigenstates of the self-Hamiltonian $`\mathrm{\Omega }\sigma _x`$) are pointer states . It should be noted that $`\overline{dP}_0`$ or $`\overline{dF}_0`$ for, say, $`y=\pm 1`$ states ($`\sigma _y`$-eigenstates) is only $`50\%`$ worse than for the pointers, for reasons that are specific to our small system.
Pointer states are the easiest to find out. โ Given the assumptions of the argument above, we have seen that pointer states do not depend on the kind of measurement carried out by the observer on the environment or on its efficiency. This robustness of pointer states might convey the wrong impression that all kinds of measurements are equivalent from the point of view of the observer trying to find out about the system by monitoring its environment. In what follows we give two examples which strongly suggest that the measurement of the environment states correlated with the pointer basis of the system is the most efficient one in gaining information about the state of the system.
We begin with the two-level atom. In the limit of $`\mathrm{\Omega }1`$, the pointer states are eigenstates of the driving self-Hamiltonian $`\mathrm{\Omega }\sigma _x`$. For $`\eta =0`$ the UME has a stationary mixed state $`\rho _s=I/2+O(1/\mathrm{\Omega })`$. Suppose that we start monitoring the environment of the atom at $`t=0`$ (detectors are turned on at $`t=0`$, and $`\eta (t)=\eta \theta (t)`$). How fast do we find out about $`๐ฎ`$? This can be measured by the purity of the conditional state. For $`\eta 1`$ the response of $`\rho `$ to the switching-on of $`\eta `$ at $`t=0`$ can be described by a small perturbation of the density matrix $`\delta \rho =[\sigma _x\delta x+\sigma _y\delta y+\sigma _z\delta z]/2`$ such that $`\rho \rho _s+\delta \rho `$. The evolution of $`\delta \rho `$ is described by
$`d\delta x`$ $`=`$ $`dt(\delta x/2)+dn\left({\displaystyle \frac{2R\mathrm{cos}\varphi }{1+2R^2}}\right),`$ (9)
$`d\delta y`$ $`=`$ $`dt(\delta y/22\mathrm{\Omega }\delta z)+dn\left({\displaystyle \frac{2R\mathrm{sin}\varphi }{1+2R^2}}\right),`$ (10)
$`d\delta z`$ $`=`$ $`dt(\delta z+2\mathrm{\Omega }\delta y)+dn\left({\displaystyle \frac{1}{1+2R^2}}\right),`$ (11)
where $`dn=dN\overline{dN}`$ and $`\overline{dN}=\eta dt(R^2+1/2)`$. For $`R1`$ a formal solution of these stochastic differential equations leads to a noise-averaged purity
$`\overline{P}(t)\mathrm{Tr}(\rho _s^2)+\mathrm{Tr}(\overline{\delta \rho ^2})={\displaystyle \frac{1}{2}}+{\displaystyle \frac{1}{2}}\overline{[\delta x_t^2+\delta y_t^2+\delta z_t^2]}=`$ (12)
$`{\displaystyle \frac{1}{2}}+\eta {\displaystyle \frac{R^2\mathrm{cos}^2\varphi }{1+2R^2}}(1e^t)+\eta {\displaystyle \frac{1+4R^2\mathrm{sin}^2\varphi }{6(1+2R^2)}}(1e^{3t/2}).`$ (13)
For any time $`t>0`$ the highest purity is obtained for homodyne ($`R1`$) measurement of $`\sigma _x`$ ($`\varphi =0`$). As anticipated, this is the measurement in the basis of environmental states correlated with the pointer states of the system. The purity saturates for $`t1`$ at
$$P_{\mathrm{}}=\frac{1}{2}+\frac{\eta }{6}\left(3\mathrm{cos}^2\varphi +2\mathrm{sin}^2\varphi \right),$$
(14)
for $`R1`$. The small $`\eta `$ measurements in the pointer state $`x`$basis ($`\varphi =0`$) are only $`50\%`$ better than in the $`y`$basis ($`\varphi =\pi /2`$), (see Fig.1). As mentioned before, in the two-level atom pointer states are not well distinguished from the chaff by the predictability sieve.
To try with an example known for well distinguished pointer states let us pick the quantum Brownian motion at zero temperature. We can think of the environment quanta as phonons. The CME obtains from Eqs.(1,3) by a formal replacement $`ca`$, where $`a,a^{}`$ are bosonic annihilation/creation operators,
$`d\rho `$ $`=`$ $`dt\left(a\rho a^{}{\displaystyle \frac{1}{2}}a^{}a\rho {\displaystyle \frac{1}{2}}\rho a^{}a\right)`$ (16)
$`+(dN\overline{dN})\left({\displaystyle \frac{(a+\gamma )\rho (a^{}+\gamma ^{})}{\mathrm{Tr}[(a+\gamma )\rho (a^{}+\gamma ^{})]}}\rho \right).`$
This equation is valid in the rotating wave approximation. After the replacement $`ca`$ in Eq.(4), we get
$$\overline{dN}=\eta dt[R^2+aRe^{i\varphi }+a^{}Re^{+i\varphi }+a^{}a]$$
(17)
For $`R=0`$ the measurement drives the conditional state to the ground state of the harmonic oscillator. In the homodyne limit ($`R\mathrm{}`$) the phonodetector current gives information about the coherent amplitude $`a`$ of the state. The second line of Eq.(16) vanishes for pure coherent states; the conditional state tends to be localized around coherent states.
For pointer states fidelity loss $`\overline{dF}_0=\mathrm{Tr}[\rho _0d\rho _0^{\mathrm{UME}}]=\mathrm{Tr}[\rho _0a^{}\rho _0a\rho _0a^{}a]`$ is the least. It vanishes if $`\rho _0`$ is a coherent state, $`\rho _0=|zz|`$ ($`a|z=z|z`$): coherent states are perfect pointers. In contrast to other states like, say, number eigenstates initially they do not lose either purity ($`\overline{dP}_0=0`$) or fidelity ($`\overline{dF}_0=0`$) (but see ).
We expect homodyne measurement to be provide more information than phonodetection. To support this, pick a coherent state $`|z`$. If $`r=|z|1`$, then $`+z|z0`$ and $`a^{}|zz^{}|z`$. In this approximation, a general density matrix in the subspace spanned by $`|\pm z`$ is
$`\rho `$ $`=`$ $`{\displaystyle \frac{1+A}{2}}|+z+z|+{\displaystyle \frac{1A}{2}}|zz|+`$ (19)
$`+C|+zz|+C^{}|z+z|.`$
Here $`A[1,+1]`$. Substitution of Eq.(19) into Eq.(16), and subsequent left and right projections on $`|\pm z`$, give stochastic differential equations for $`A`$ and $`C`$. These equations are most interesting in two limits. In the phonodetection limit ($`R=0`$) they are $`dA=0`$, $`dC=C[2r^2dt+(dN\overline{dN})]`$. The off-diagonal $`C`$ decays after the decoherence time of $`1/r^21`$. $`A`$ does not change, phonodetection does not produce any purity. Phonodetection is a very poor choice: by this measurement we learn nothing about the system! In the opposite homodyne detection limit ($`R\mathrm{}`$) the noise $`dN\overline{dN}`$ can be replaced (up to a constant) by a white-noise $`dW`$ such that $`\overline{dW}=0`$ and $`\overline{dW^2}=dt`$ . Again, $`C`$ decays after the decoherence time of $`1/r^2`$. Introducing $`B=\mathrm{tanh}^1(A)`$, defining a time scale $`\tau 4t\eta r^2\mathrm{cos}^2(\varphi \theta )`$ (here $`\theta `$ is the phase of the coherent state, $`z=r\mathrm{exp}(i\theta )`$), and a noise $`d\zeta 2\sqrt{\eta }r\mathrm{cos}(\varphi \theta )dW`$ ($`d\zeta ^2=d\tau `$), we get the following Stratonovich stochastic equation
$$\frac{dB}{d\tau }=\mathrm{tanh}B+\frac{d\zeta }{d\tau }.$$
(20)
Suppose that at $`t=0`$ we had $`A=B=0`$ and $`C=0`$. This is the most mixed state possible in our subspace. The probability distribution for $`A`$ at time $`\tau >0`$ is
$$p(\tau ,A)=\frac{(2\pi \tau )^{1/2}}{(1A^2)^{3/2}}\mathrm{exp}\left(\frac{\tau }{2}\frac{\mathrm{ln}^2\left(\frac{1+A}{1A}\right)}{8\tau }\right)$$
(21)
This distribution is localized at $`A=0`$ for $`\tau =0`$ but after a time scale $`\tau 1`$ it becomes concentrated at $`A=\pm 1`$ (see Fig.2). By these times the conditional state is almost certainly one of the coherent states $`|\pm z`$ and purity is $`1`$. The asymptotic bimodal distribution is obtained the fastest for a homodyne tuned to the phase of the coherent states, $`\varphi =\theta `$. This result is in sharp contrast to the nil result for phonodetection. In Fig.3 we plot three realizations of a stochastic trajectory $`A(\tau )`$.
For any $`\eta <1`$ purity becomes $`1`$ after a time proportional to $`1/\eta `$. A patient observer gets full information about the system monitoring only a small part of the environment: information about pointer states is recorded by the environment in a redundant way .
In the above example we assumed that $`r=|z|1`$ so that $`+z|z0`$ and $`a^{}|zz^{}|z`$. This convenient assumption also naturally separates the decoherence and purification timescales ($`1/r^2`$) from the timescale for decay towards the ground state ($`1`$). On the fast timescales $`1/r^2`$ we can neglect the decay and that is why our system remains in the $`|\pm z`$-subspace. In this sense our calculation is self-consistent.
Concluding remarks. โ The aim of our paper was to study the issue of the prefered states in the context of conditional master equations using the predictability sieve. We have shown under reasonable, but not completely general conditions, that the most classical states of a system which is being monitored are independent both of the type of measurement and of the detector efficiency. Furthermore, we have found indications that the best measurements of the environment for gaining information about a system extract data about its pointer basis.
Acknowledgements. We are grateful to H.M.Wiseman for discussions and critical comments. This research was supported in part by NSA.
|
warning/0006/nucl-ex0006001.html
|
ar5iv
|
text
|
# New Limit on the ๐ท Coefficient in Polarized Neutron Decay
## I Introduction
$`CP`$ violation has been observed so far only in the decays of neutral kaons . Recently evidence for the implied $`T`$ violation in the neutral kaon system has been reported . These effects could be due to the Kobayashi-Maskawa phase in the Standard Model . However, these observations could also be due to new physics, and it is well-established that new sources of $`CP`$ violation are required by the observed baryon asymmetry of the universe . Many extensions of the Standard Model contain new sources of $`CP`$ violation and can be probed in observables for which the contribution of the Kobayashi-Maskawa phase in the Standard Model is small. The present experiment searches for $`CP`$ violation in one such observable, a $`T`$-odd correlation in the decay of free neutrons.
The differential decay rate for a free neutron can be written
$`dWS(E_e)dE_ed\mathrm{\Omega }_ed\mathrm{\Omega }_\nu [1+a{\displaystyle \frac{๐ฉ_e๐ฉ_\nu }{E_eE_\nu }}`$ (1)
$`+{\displaystyle \frac{๐}{J}}(A{\displaystyle \frac{๐ฉ_e}{E_e}}+B{\displaystyle \frac{๐ฉ_\nu }{E_\nu }}+D{\displaystyle \frac{๐ฉ_e\times ๐ฉ_\nu }{E_eE_\nu }})],`$ (2)
where $`p_e`$, $`E_e`$ and $`p_\nu `$, $`E_\nu `$ are the momentum and energy of the outgoing electron and neutrino, respectively, $`S(E_e)`$ is a phase space factor, and $`๐`$ is the neutron spin. The triple-correlation $`D๐(๐ฉ_e\times ๐ฉ_\nu )`$ is odd under motion reversal, and can be used to measure time reversal invariance violation when final state interactions are taken into account. Note that in the rest frame of the neutron, conservation of momentum allows the transformation of the triple-correlation term into
$$D\frac{๐}{J}\frac{๐ฉ_e\times ๐ฉ_p}{E_eE_\nu }$$
where $`๐ฉ_p`$ is the momentum of the recoil proton.
<sup>1</sup><sup>1</sup>footnotetext: Present address: National Institute of Standards and Technology, Gaithersburg, Maryland<sup>2</sup><sup>2</sup>footnotetext: Present address: National Central University, Chung-Li, Taiwan<sup>3</sup><sup>3</sup>footnotetext: Present address: Hamilton College, Clinton, NY<sup>4</sup><sup>4</sup>footnotetext: Present address: SAIC, Somerville, MA<sup>5</sup><sup>5</sup>footnotetext: Present address: Personify, Inc., San Francisco, CA
The $`D`$ coefficient is sensitive only to $`T`$-odd interactions with vector and axial vector currents. In a theory with such currents, the coefficients of the correlations depend on the magnitude and phase of $`\lambda =|\lambda |e^{i\varphi }`$, where $`|\lambda |=|g_A/g_V|`$ is the magnitude the ratio of the axial vector to vector form factors of the nucleon. In this notation, the coefficients are given by
$`a={\displaystyle \frac{1|\lambda |^2}{1+3|\lambda |^2}},A=2{\displaystyle \frac{|\lambda |\mathrm{cos}\varphi +|\lambda |^2}{1+3|\lambda |^2}},`$ (3)
$`B=2{\displaystyle \frac{|\lambda |\mathrm{cos}\varphi |\lambda |^2}{1+3|\lambda |^2}},D=2{\displaystyle \frac{|\lambda |\mathrm{sin}\varphi }{1+3|\lambda |^2}}.`$ (4)
The most accurate determinations of $`|\lambda |`$ (current world average $`|\lambda |=1.2670\pm 0.0035)`$ come from measurements of $`A`$ . The coefficients $`a`$, $`A`$, and $`B`$, respectively, are measured to be $`0.102\pm 0.005`$, $`0.1162\pm 0.0013`$, and $`0.983\pm 0.004`$ . Several previous experiments found the value of $`D`$, and thus $`\mathrm{sin}\varphi `$, to be consistent with zero at a level of precision well below 1%. The three most recent such measurements found $`D=(1.1\pm 1.7)\times 10^3`$ and $`D=(2.2\pm 3.0)\times 10^3`$ , and $`D=(2.7\pm 5.0)\times 10^3`$ , constraining $`\varphi `$ to $`180.07^{}\pm 0.18^{}`$ .
Final state interactions give rise to phase shifts of the outgoing electron and proton Coulomb waves that are time reversal invariant but motion reversal non-invariant. Thus $`D`$ has terms that arise from phase shifts due to pure Coulomb and weak magnetism scattering. The Coulomb term vanishes in lowest order in V-A theory , but scalar and tensor interactions could contribute. The Fierz interference coefficient measurements can be used in limiting this possible contribution to
$$|D^{EM}|<(2.8\times 10^5)\frac{m_e}{p_e}.$$
(5)
Interference between Coloumb scattering amplitudes and the weak magnetism amplitudes produces a final state effect of order ($`E_{e}^{}{}_{}{}^{2}/p_em_n`$). This weak magnetism effect is predicted to be
$$D^{WM}=1.1\times 10^5.$$
(6)
The $`D`$ coefficient has also been measured for <sup>19</sup>Ne decay, with the most precise experiment finding $`D_{Ne}=(4\pm 8)\times 10^4`$ . The predicted final state effects for <sup>19</sup>Ne are approximately an order of magnitude larger than those for the neutron and may be measured in the next generation of <sup>19</sup>Ne experiments. For <sup>8</sup>Li, a triple-correlation of nuclear spin, electron spin and electron momentum has been measured, with the most precise measurement at $`R=(0.9\pm 2.2)\times 10^3`$ . Unlike $`D`$, a nonzero $`R`$ requires the presence of scalar or tensor couplings and thus is a tool to search for such couplings. The electric dipole moments (EDMs) of the electron , neutron , and <sup>199</sup>Hg atom are arguably the most precisely-measured $`T`$-violating parameters and bear on many of the same theories as $`D`$. Table I summarizes the current constraints on $`D`$ from analyses of data on other $`T`$-odd observables for the Standard Model and extensions . For lines 2-5 these limits are derived from the measured neutron or <sup>199</sup>Hg EDM.
In the nearly two orders of magnitude between the present limit on $`D`$ and the final state effects lies the opportunity to directly observe or limit new physics. Moreover, accurate calculations of magnitude and energy dependence of the final state effects can be made to extend the range of exploration still further.
## II Overview of the emiT Detector
In the emiT apparatus, a beam of cold neutrons is polarized and collimated before it passes through a detection chamber with electron and proton detectors (four each). A schematic of the experiment is shown in Figure 1.
The most significant improvements over previous experiments are the achievement of near-unity polarization ($`>93\%`$ compared to 70% in ) and the construction of a detector with greater acceptance and greater sensitivity to the $`D`$ coefficient. The octagonal arrangement of the eight detector segments gives them nearly full coverage of the 2$`\pi `$ of azimuthal angle around the beam, nearly twice the angular acceptance in previous experiments, and the detector segments are longer than in previous experiments. The placement of the two types of detectors at relative angles of 135 is also an improvement over previous experiments, in which the coincidences were detected at 90. While the cross product is greatest at 90, the preference for larger electron-proton angles in the decay makes placement of the detectors at 135 the best choice to achieve greater symmetry, greater acceptance, and greater sensitivity to $`D`$ (see Figure 2).
Combined with the higher neutron polarization from the supermirror polarizer our geometry provides for an overall sensitivity to $`D`$ that is a factor of $`7`$ greater than previous measurements, assuming the same cold neutron beam flux.
The first run of the experiment was conducted at the NIST Center for Neutron Research (NCNR) in Gaithersburg, MD. The experimental apparatus is outlined below, while more detailed descriptions can be found in .
### A Polarized Neutron Beam
The NCNR operates a 20-MW, heavy-water-moderated research reactor. Neutrons from the reactor pass through a liquid hydrogen moderator to make cold neutrons with an approximately Maxwellian velocity distribution at a temperature of about 40 K. The average neutron velocity is about 800 m/s. The neutrons are transported 68 meters to the apparatus via a <sup>58</sup>Ni-lined neutron guide. Neutrons are totally internally reflected if they enter with an angle of incidence less than 2 mrad for each ร
of de Broglie wavelength. The capture flux of the neutrons was measured using a gold foil activation technique to be $`\rho _nv_0=1.4\times 10^9`$n/cm<sup>2</sup>/s (where $`v_0`$ = 2200 m/s) at the end of the neutron guide. (The capture flux quantifies the neutron density in the detector for the polychromatic beam.) The beam passes through a cryogenic beam filter of 10-15 cm of single crystal bismuth which filters out residual fast neutrons and gamma-rays.
The neutrons are polarized with a double-sided bender-type supermirror polarizer obtained from the Institut Laue-Langevin in Grenoble, France . The supermirror consists of 40 Pyrex plates coated on both sides with cobalt, titanium, and gadolinium layers that maximize the reflection of neutrons with the desired spin state while absorbing nearly all neutrons of the opposite spin state. The supermirror was measured to polarize a 4.5 cm by 5.5 cm beam with 24% transmission relative to the incident unpolarized flux. The neutron polarization was determined to be $`>93\%`$ ($`95\%`$ CL).
The neutrons travel the one meter from the polarizer to the spin-flipper inside a Be-coated glass flight tube in which a small helium overpressure is maintained to minimize beam attenuation via air scattering. The neutrons, which have spins that are transverse to their motion, then pass through two layers of aluminum wires which comprise the current-sheet spin flipper. When the current in the second layer is antiparallel to that in the first there is no net magnetic field and the neutron polarization is unaffected. When the currents are parallel, the neutron spin does not adiabatically follow the rapid change in field orientation and thus the sense of $`๐๐`$ is reversed. Downstream of the spin flipper, weak magnetic fields adiabatically rotate the spin to longitudinal, i.e. parallel or antiparallel to the neutron momentum. The longitudinal guide fields are 2.5 mT upstream and 0.5 mT within the detector. Figure 3 shows the spin transport system. The polarization direction is reversed every 5 seconds.
In the detection region, the longitudinal field is produced by eight 50 amp-turn current loops of 1 m diameter. The loops are aligned to within 10 mrad of the detector axis using a sensitive field probe and an AC lock technique. Additional coils canceled the transverse components of the Earthโs field and local gradients of 7.5 $`\mu `$T/m.
The vacuum chamber begins at the spin flipper with two meters of Be-coated flight tubes, through which the neutrons travel toward the collimator region. Two collimators of 6 cm and 5 cm diameter openings separated by 2 m define the beam. These and 5 additional โscrapersโ between them consist of rings of <sup>6</sup>LiF which absorb the neutrons. Behind each ring is a thick ring of high-purity lead which absorbs the gamma-rays from the reactor and those produced by neutron captures upstream. Between scrapers, the walls of the beam tube are lined with <sup>6</sup>Li-loaded glass to absorb stray neutrons.
A fission chamber mounted behind a sheet of <sup>6</sup>Li-glass with a 1 mm pinhole aperture was scanned across the beam to obtain a cross-sectional profile of the intensity as shown in Figure 4. The neutron intensity was measured before and after the experiment. To determine the polarization at the entrance to the detector, the beam passed through a second, single-sided, analyzing supermirror directly in front of the scanning detector, and the ratio of intensities with the spin flipped and unflipped was measured. The resulting flippng ratio measures a combination of the neutron-spin-dependent transmission efficiencies of the two supermirrors and the neutron spin flipping eficiency. From this, and assumptions about the spin flipping efficiency, we can determine the product of polarization efficiencies for the two supermirrors (polarizer and analyzer). When the upper limit of 100% spin flipping efficiency is used, a lower limit of the neutron beam polarization of $`93\%`$ ($`95\%`$ CL) is found. This lower limit also includes the assumption that the flipping ratio for a pair of supermirrors identical to our analyzer would be less than that of a pair of supermirrors identical to our polarizer by a factor of $`2\pm 0.5\%`$.
Downstream of the detection region the vacuum chamber diameter increases to 40.6 cm, terminating with a <sup>6</sup>Li-glass beam stop 2.8 m from the end of the detector. A 1 mm diameter pinhole at the center of the beamstop allows about 1% of the beam to pass through a silicon window into a fission chamber detector that continuously monitors the neutron flux.
### B Detector System
Eight detectors surround the beam, each 10 cm from the beam axis as shown in Figure 5. The octagonal geometry places electron and proton detectors at relative angles of 45 and 135 degrees. Coincidences are counted between detectors at relative angles of 135 degrees.
#### 1 Electron Detectors
The electron detectors are slabs (8.4 cm x 50 cm x 0.64 cm) of BC408 plastic scintillator connected on each end to curved lucite light-guides that channel the light to Burle 8850 photomultiplier tubes. Each photomultiplier tube is surrounded by a mu-metal magnetic shield and a pair of nested solenoids acting as an active magnetic shield. This combination of active and passive magnetic shielding had a factor of 10 less impact (0.5 $`\mu `$T) on the guide field at the beam center than the mu-metal alone.
The scintillator thickness of 0.64 cm is just greater than that necessary to stop the most energetic (782 keV) of the electrons from neutron. The scintillators are wrapped with aluminized mylar and aluminum foil to prevent charging and to shield the detectors from x-rays and field-emission electrons in the vacuum chamber. For each segment, the energy response was calibrated with cosmic-ray muons and conversion electrons from <sup>207</sup>Bi and <sup>113</sup>Sn (see Figure 6.)
#### 2 Proton Detectors
Each proton detector has an array of 12 PIN diodes of 500 micrometer thickness arranged in two rows of 6. The diodes are held within a stainless steel high voltage electrode. Over each diode an open cylinder protrudes from the face of the electrode, shaping the field to focus and accelerate the protons as shown in Figure 7. Thus each diode collects protons focused from a region 4 cm $`\times `$ 4 cm even though it has an active area of only 1.8 cm $`\times `$ 1.8 cm.
The diodes and their electronics are held at -30 to -40 kV. Between the electrode and the beam is a frame strung with 80 0.08-mm gold-plated tungsten wires that define a plane of electrical ground. Protons drift in a field-free region until they pass this plane, and then are accelerated by the high voltage and focused onto the nearest PIN diode below. Near both ends of the detector array are two cryopanels held at liquid nitrogen temperature. Water vapor, released predominantly by the scintillators and other plastic components, is pumped onto the cryopanels to prevent condensation on the cooled PIN diodes.
The charge in the PIN diode produced by each proton is amplified by 10 V/pC with a preamplifier mounted directly behind the PIN diode. These circuits and the PIN diodes are cooled with liquid nitrogen to about 0C to decrease electronic noise. Preamplifier signals are processed in a custom VME-format shaper/ADC board with programmable gain and operating mode parameters. The PIN diodes were calibrated with x-rays from an <sup>241</sup>Am source as shown in Figure 8.
#### 3 Background
The background in the detectors was primarily related to the beam or to the high voltage bias. Closing the beam shutter upstream of the neutron filter stops virtually all neutrons and about 1/3 of the gamma-rays coming from the reactor along the beamline. With the shutter closed, the rates in each detector were less than 100 Hz, primarily from dark current, reactor gamma-rays, and cosmic rays. With the shutter open, the detectors see an increased gamma-ray flux primarily from neutron captures in the apparatus, triggering the detectors at less than 1 kHz per electron detector and less than 1 kHz for all PIN diodes combined. This results in deadtime less than 3% for the beam-related background. At its worst, the high-voltage-related background, consisting of x-rays, light, electrons, and ions, led to rates in the hundreds of kHz in the detectors. It was reduced at times by conditioning and cleaning of electrodes but varied by orders of magnitude during the run.
### C Data Acquisition
A block diagram of the data acquisition system is shown in Figure 9. The identification of neutron decay events is simplified by the fact that the proton signal is observed 0.5 $`\mu `$s to 2 $`\mu `$s after the electron signal. The recoil protons, with maximum energies of only 750 eV, require this time to drift from the point of decay to the face of the proton detector. Events are accepted by the coincidence trigger when the electron signal arrives within a coincidence time window $`\pm \tau _{coinc}/2`$ of a proton signal. The duration $`\tau _{coinc}`$ of this window was originally 14 $`\mu `$s and was shortened to 7 $`\mu `$s midway through the experiment to reduce the system deadtime. Each stored event contains the location (PIN diode) and energy for the proton event, location (electron detector), energy of the electron event, relative time between individual signals from the two phototubes in the electron detector, relative time of arrival of the proton and electron signals, and the orientation of the neutron polarization. Every 30 seconds during the data collection, information is recorded from the system monitors which include system livetime, magnet currents, neutron flux at the beam stop, vacuum pressure, proton detector high voltage, and high voltage leakage current. Periodically, the data acquisition collects singles spectra from all of the individual detectors.
## III Experimental Run
### A Data Collection
The experiment was installed at the NCNR during December 1996 and January 1997. From February through August 1997, 50 GB of data were collected and stored. The data are divided into 626 files representing continuous runs, typically four hours in duration. These are grouped into 125 series, within which running conditions varied little. For one week in August a systematic test was run in which the beam was distorted and the polarization guide field direction changed. The purpose and results of this test will be described in Section IV.
Instabilities in the proton detector high voltage made it impossible to operate all channels of the detectors at all times. Sometimes the electrodes simply would not hold the necessary voltage, and at other times a large spark or series of sparks would damage the electronics held at high voltage. Less than half of the data were collected when all four proton detectors were functioning. Another limitation to the detector uniformity were variations in the measured proton energy deposited in the PIN diodes. In preliminary tests, the surface dead-layers of the PINs were measured to be 20$`\pm 2`$ $`\mu `$g/cm<sup>2</sup> as specified by the manufacturer, Hamamatsu. In a dead-layer of this thickness a 35 keV proton loses 10 keV of energy. The proton energies measured during the experiment, however, were 12-18 keV, an average of 20 keV below the energy imparted to them by acceleration through 34-38 kV (see Figure 10.) With widths (FWHM) of approximately 10 keV, these peaks are not well separated from the background.
High background rates necessitated the setting of thresholds at levels such that some neutron decay events were also rejected. This and the data acquisition deadtime were the primary limitations to the statistics of the experiment. A deadtime per event of 2 ms was necessary for stability of the system. Even with the reduction in length of the coincidence window, the high rate of background kept the system at 40-60% deadtime for most of the data collection period.
### B Event Selection
Figure 11 shows an example of the relative time spectrum for the coincidence data. The large center spike, originating mainly from multiple gamma rays produced by neutron captures in the apparatus, defines zero time difference. The neutron decay events are accepted within a window 0.35 $`\mu `$s to 0.9 $`\mu `$s after the prompt peak. This window contains the majority of the neutron decay protons, while excluding the tail of the prompt peak and the low-signal-to-background tail of the proton peak. The background to be subtracted from these events is estimated using the rates in regions to either side of the decay and zero-time peaks.
Events are also selected on the basis of measured proton energy to reduce the amount of background to be subtracted. The energy range accepted is chosen solely by minimizing the fractional statistical uncertainty in the number of neutron decay events for each PIN diode-electron detector pair. Specifically, if $`N_\mathrm{\Delta }`$ is the number of coincidences counted by subtracting the background from the coincidences in the 0.35 to 0.9 $`\mu `$s window, the energy range is chosen to minimize
$$\frac{\sigma _{N_\mathrm{\Delta }}}{N_\mathrm{\Delta }}\frac{\sqrt{1+1/f}}{\sqrt{N_\mathrm{\Delta }}},$$
(7)
where $`f`$ is the signal to background ratio in this energy range. This increases the overall signal to background on the 15 million good events from 0.8 to 2.5.
## IV Data Analysis and Uncertainty Estimation
### A Determination of $`D`$ from Coincidence Events
For each PIN diode-electron detector pair in a given data series, the count rate can be expressed as
$`N_\pm ^{\alpha i}=N_0ฯต^\alpha ฯต^i[K_1^{\alpha i}+aK_a^{\alpha i}\pm P\widehat{\sigma }`$ (8)
$`(A๐_A^{\alpha i}+B๐_B^{\alpha i}+D๐_D^{\alpha i})],`$ (9)
where $`N_0`$ is a constant proportional to the beam flux, the $`ฯต^\alpha `$ and $`ฯต^i`$ are detector efficiencies for a PIN diode and electron detector respectively. The average of the neutron polarization vector over the detector volume, given by $`P\widehat{\sigma }`$,is assumed to be uniform and constant over time, lying along the direction of the 0.5 mT guide field. The $`\pm `$ signs correspond to the two signs of the polarization. The factors $`K_1^{\alpha i}`$ and $`K_a^{\alpha i}`$ are geometric factors derived from Equation 2 by integrating $`1`$ and $`๐ฉ_e๐ฉ_\nu /E_eE_\nu `$, respectively, over the beta-decay phase space, the neutron beam volume, and the acceptance of each electron-detectorโPIN-diode detector pair. Similarly, the factors $`๐_A^{\alpha i}`$, $`๐_B^{\alpha i}`$, and $`๐_D^{\alpha i}`$, are obtained by integrating the vectors: $`๐ฉ_e/E_e`$, $`๐ฉ_\nu /E_\nu `$, and $`(๐ฉ_e\times ๐ฉ_\nu )/E_eE_\nu `$.
We produce the following efficiency-independent asymmetries
$$w^{\alpha i}=\frac{N_+^{\alpha i}N_{}^{\alpha i}}{N_+^{\alpha i}+N_{}^{\alpha i}}.$$
(10)
From Equation 9 we get
$$w^{\alpha i}=P\widehat{\sigma }(A\stackrel{~}{๐}_A^{\alpha i}+B\stackrel{~}{๐}_B^{\alpha i}+D\stackrel{~}{๐}_D^{\alpha i}),$$
(11)
where we use the definitions
$$\stackrel{~}{๐}_A^{\alpha i}=\frac{๐_A^{\alpha i}}{K_1^{\alpha i}+aK_a^{\alpha i}}\mathrm{etc}.$$
(12)
Consider the two detector pairings PIN<sub>a</sub>-E<sub>1</sub> and PIN<sub>b</sub>-E<sub>2</sub> indicated in Figure 12. The corresponding values of $`๐_D^{\alpha i}`$ have opposite sign while $`๐_A^{\alpha i}`$ and $`๐_B^{\alpha i}`$ have the same sign. We therefore combine asymmetries from two proton-electron detector pairings to produce the combination
$`v^{b2:a1}={\displaystyle \frac{1}{2}}[w^{b2}w^{a1}]`$ (13)
$`={\displaystyle \frac{1}{2}}P\widehat{\sigma }[D(\stackrel{~}{๐}_D^{b2}\stackrel{~}{๐}_D^{a1})`$ (14)
$`+A(\stackrel{~}{๐}_A^{b2}\stackrel{~}{๐}_A^{a1})+B(\stackrel{~}{๐}_B^{b2}\stackrel{~}{๐}_B^{a1})].`$ (15)
For uniform detection efficiency the difference $`(\stackrel{~}{๐}_D^{b2}\stackrel{~}{๐}_D^{a1})`$ lies along the detector axis, $`\widehat{z}`$, while the differences $`(\stackrel{~}{๐}_A^{b2}\stackrel{~}{๐}_A^{a1})`$ and $`(\stackrel{~}{๐}_B^{b2}\stackrel{~}{๐}_B^{a1})`$ lie perpendicular to the detector axis. For a polarized neutron beam with perfect cylindrical symmetry aligned with the detector axis, $`\stackrel{~}{๐}_D^{b2}\widehat{z}=\stackrel{~}{๐}_D^{a1}\widehat{z}`$, and
$$v^{b2:a1}=PD\stackrel{~}{๐}_D^{b2}\widehat{z}=PD\stackrel{~}{๐}_D^{a1}\widehat{z}.$$
(16)
Departures from perfect symmetry and perfect alignment of the neutron polarization require that the $`A`$ and $`B`$ correlation terms be retained in Equation 15. The resulting systematic effects are discussed in Section IV C.
Additionally, as shown in Figure 12, there are two classes of electron-PIN pairs: those that make an angle smaller than 135 ($`b2:a1`$) or an angle larger than 135 ($`a2:b1`$).
We thus separate our data into a small-angle group and a large-angle group giving two statistically independent results for each PIN-diodeโelectron-detector pairing.
### B Monte Carlo Methods
We use two Monte Carlo calculations to determine the values of $`K_1`$, $`K_a`$, $`๐_A`$, $`๐_B`$, and $`๐_D`$. The results from these two completely independent calculations are in excellent agreement. In both calculations neutron decay events are generated randomly within a trapezoid-cylindrical geometry (i.e. a tube with divergence) that can be offset with respect to the detector axis. A realistic beam profile, representative of Figure 4, can be modeled by combining results from several different trapezoids. In one of the Monte Carlo calculations the tracking of protons and electrons is done with the CERN Library GEANT3 Monte Carlo package , while in the other tracking is implemented within the code itself. In both, the emiT detector geometry is specified with uniform efficiency over the active area of each scintillator and over the square focusing region of each PIN diode.
The constants defined in Equation 9 are given by
$$K_x^{\alpha i}=\delta ^{\alpha i}X$$
(17)
where $`X=1`$, $`๐ฉ_e๐ฉ_\nu /E_eE_\nu `$, $`๐ฉ_e/E_e`$, $`๐ฉ_\nu /E_\nu `$, and $`๐ฉ_e\times ๐ฉ_\nu /E_eE_\nu `$ for $`x=1,a,A,B,`$ and $`D`$, respectively. We have studied systematic uncertainties associated with potential non-uniformities in the beta efficiencies and included them in the final uncertainty for the constants $`K_x^{\alpha i}`$. These constants (a total of 11, taking into account the three directions for each vector) are accumulated in a file that is read to calculate the factors $`v`$ (Equation 15) for different orientations of the polarization.
Values of $`|\stackrel{~}{๐}_D^{\alpha i}\widehat{z}|`$ are used directly in the interpretation of the result for $`D`$. Variations among the PIN diode pairs of individual values of $`\stackrel{~}{๐}_D^{\alpha i}`$ within a given proton segment are negligible, and average values ($`|\stackrel{~}{๐}_D\widehat{z}|`$) can be used. They are found to be $`0.424\pm 0.010`$ and $`0.335\pm 0.020`$, for the small- and large-angle coincidences, respectively. The uncertainties are primarily from uncertainties in the geometry of the beam. Values for the other $`K_x^{\alpha i}`$ are used in the estimation of systematic uncertainties described in the following section.
### C Discussion of Systematic Uncertainties
The largest of the systematic effects can be shown to be the contributions to the $`v`$ (Equation 18) that arise due to the misalignment of the neutron polarization with respect to the detector axis. A transverse component of the polarization produces a significant contribution to $`v^{b2:a1}`$ because the vector differences $`\stackrel{~}{๐}_A^{b2}\stackrel{~}{๐}_A^{a1}`$ and $`\stackrel{~}{๐}_B^{b2}\stackrel{~}{๐}_B^{a1}`$ are predominantly perpendicular to the detector axis. (For example, $`\stackrel{~}{๐}_A^{b2}\stackrel{~}{๐}_A^{a1}`$ is proportional to the integral of $`๐ฉ_e(E_1)๐ฉ_e(E_2)`$ and is directed horizontally to the left in Figure 12. The difference $`\stackrel{~}{๐}_B^{b2}\stackrel{~}{๐}_B^{a1}`$ is antiparallel to $`\stackrel{~}{๐}_A^{b2}\stackrel{~}{๐}_A^{a1}`$.) For an azimuthally symmetric neutron beam, it can be shown that for each proton detector segment (labeled with subscripts $`\eta `$= I, II, III, IV) the weighted average of the $`v^{\alpha i:\beta j}`$ for all large or small detector pairs can be expressed as
$$v_\eta ^{l/s}=PD(\stackrel{~}{๐}_D^{l/s}\widehat{\sigma })+\alpha _\eta ^{l/s}\mathrm{sin}\theta _\sigma \mathrm{sin}(\varphi _\eta \varphi _\sigma ),$$
(18)
where $`\theta _\sigma `$ and $`\varphi _\sigma `$ are the polar and azimuthal angles of $`\widehat{\sigma }`$, and $`\varphi _\eta `$=0, 90, 180, and 270 respectively for detectors I, II, III, and IV. This dependence can be derived analytically for zero beam radius and is confirmed by Monte Carlo simulations for symmetric beams of finite radius. The coefficients $`\alpha _\eta `$ measure the combined effects of the $`A`$ and $`B`$ correlations for each proton detector segment.
If the symmetry of the four sets of proton detectors were perfect, i.e. $`\alpha _\mathrm{I}=\alpha _{\mathrm{II}}=\alpha _{\mathrm{III}}=\alpha _{\mathrm{IV}}`$, the contributions due to the $`A`$ and $`B`$ coefficients would average to zero, and Equation 16 would be valid, even with a polarization misalignment. In the absence of perfect symmetry, these contributions do not cancel when the four proton detectors are combined, and a false $`D`$ contribution would result from the application of Equation 16. This false $`D`$ is proportional to the product of two effects that are both small: the misalignment of the neutron polarization with respect to the detector axis ($`\theta _\sigma `$) and the departure from perfect symmetry of the proton detectors ($`\mathrm{\Delta }\alpha =1/2(\alpha _\mathrm{I}\alpha _{\mathrm{III}})+1/2(\alpha _{\mathrm{II}}\alpha _{\mathrm{IV}})`$). Such an effect is called the โtilting asymmetric transverse polarizationโ effect, or โTilt ATPโ.
The ATP effect was intentionally amplified for a systematic test, run with transverse polarization ($`\theta _\sigma =90^{}`$ $`\varphi _\sigma =\varphi _{\mathrm{IV}}=270^{}`$) and a distorted neutron beam. The neutron beam was distorted by blocking half of the beam with a neutron absorber placed upstream near the spin flipper. The results of this test are shown in Figure 13.
This demonstration that the experiment can measure an asymmetry consistent with the Monte Carlo calculation serves as a strong check on both the operation of the detector and the validity of the analysis method.
A false $`D`$ also arises if the polarization has transverse components not described by a simple tilt. The form of Equation 18 shows that a net azimuthal component of $`\widehat{\sigma }`$ also results in a contribution to $`v_\eta `$ that does not average to zero when data from proton segments I-IV are combined. This effect, referred to as a โtwisting asymmetric transverse polarizationโ (โTwist ATPโ) is shown by Monte Carlo simulations to be less than $`10^4`$ for azimuthal polarizations of less than 1 mrad. For this reason, all sources of guide field distortion are kept to less than 1 mrad, and materials of low magnetic permeability (less than 0.005 $`\mu _0`$) were used in the detection region. There are exceptions to this requirement, however the net effect of all additional permeability was measured to produce less than 1 mrad of distortion of the guide field anywhere in the detector region.
Variations in the neutron flux ($`\mathrm{\Phi }`$) and polarization ($`P`$) that depend on neutron helicity yield a false $`D`$. For this experiment the effects due to misalignment of the neutron spin are small, so that these systematic effects, to first order in $`\mathrm{\Delta }\mathrm{\Phi }/\overline{\mathrm{\Phi }}`$ and $`\mathrm{\Delta }P`$ are
$$D_{false}(\mathrm{\Delta }\mathrm{\Phi })=\frac{\mathrm{\Delta }\mathrm{\Phi }}{\overline{\mathrm{\Phi }}}PD\widehat{\sigma }(A\stackrel{~}{๐}_A+B\stackrel{~}{๐}_B),$$
(19)
and
$$D_{false}(\mathrm{\Delta }P)=\mathrm{\Delta }PD\widehat{\sigma }(A\stackrel{~}{๐}_A+B\stackrel{~}{๐}_B).$$
(20)
Here $`\mathrm{\Delta }\mathrm{\Phi }=\mathrm{\Phi }_{}\mathrm{\Phi }_{}`$, and $`\mathrm{\Delta }P=P_+P_{}`$. ($`P`$ in Equations 15 and 16 would be replaced by $`\overline{P}=(1/2)[P_++P_{}]`$.) The $`\stackrel{~}{๐}_A`$ and $`\stackrel{~}{๐}_B`$ are average values for all PIN-diodeโelectron-detector pairings.
Our data provide an upper limit of 0.002 for $`P\widehat{\sigma }(A\stackrel{~}{๐}_A+B\stackrel{~}{๐}_B)`$. We combine this with neutron flux monitor data for $`\mathrm{\Delta }\mathrm{\Phi }/\mathrm{\Phi }<0.004`$, concluding that $`D_{false}(\mathrm{\Delta }\mathrm{\Phi })<8\times 10^6D`$. The flipping ratio measurement has been used to derive a lower limit on the spin flipper efficiency of 82% so that $`\mathrm{\Delta }P<0.2`$, and $`D_{false}(\mathrm{\Delta }P)<4\times 10^4D`$. We conclude that both effects are negligible in this measurement.
### D Results
A final value of $`D_\eta ^{l/s}=v_\eta /(P\stackrel{~}{๐}_D^{l/s}\widehat{z})`$ is found separately for large angle and small angle pairings of each proton segment. The quantities $`v_\eta `$ are the weighted averages of all PIN-electron detector pairs, $`v^{l/s}(\alpha i:\beta j)`$, within each proton detector segment. Use of the weighted averages is justified because the systematic uncertainties described above have negligible variations among the PIN diode pairs in a given detector. The individual proton segment data ($`v_\eta `$) are then combined in an arithmetic average so that the sinusoidal variation given in Equation 16 cancels to first order in misalignments, i.e.
$`{\displaystyle \underset{\eta =\mathrm{I}}{\overset{\mathrm{IV}}{}}}v_\eta ^{l/s}=4D^{l/s}(P\stackrel{~}{๐}_D^{l/s}\widehat{z})+๐ช(\theta _\sigma \mathrm{\Delta }\alpha ).`$ (21)
The error for $`D^{l/s}`$ includes the uncertainty in the values of $`\stackrel{~}{๐}_D^{l/s}\widehat{z}`$.
$`\sigma _{D^{l/s}}^2={\displaystyle \frac{1}{4P\stackrel{~}{๐}^{l/s}\widehat{z}}}{\displaystyle \underset{\eta =\mathrm{I}}{\overset{\mathrm{IV}}{}}}\sigma _{v^{l/s}}^2+D^{l/s}\left({\displaystyle \frac{\sigma _{|\stackrel{~}{๐}_D^{l/s}\widehat{z}|}}{|\stackrel{~}{๐}_D^{l/s}\widehat{z}|}}\right)^2`$ (22)
Data for each proton segment are displayed in Figure 14, where we plot values for the eight individual $`D_\eta ^{l/s}`$.
(The sinusoidal variation is predicted by Equation 18 and also seen in the test data of Figure 13, where the amplitude is 100 times larger.)
The two independent measurements for small angle and large angle PIN-electron detector pairs can be combined in a weighted average.
$$D=\frac{D^s/\sigma _{D^s}^2+D^l/\sigma _{D^l}^2}{1/\sigma _{D^s}^2+1/\sigma _{D^l}^2}$$
(23)
The full uncertainty includes the uncertainty from the average neutron beam polarization.
$$\sigma _D^2=\left(\frac{1}{1/\sigma _{D_{(s)}}^2+1/\sigma _{D_{(l)}}^2}\right)+\left(D\frac{\sigma _{\overline{P}}}{\overline{P}}\right)^2$$
(24)
The data are also analyzed by breaking each series up into individual runs and combining PIN-electron detector pairings in the same way. The results of these analyses are consistent. The final result is $`(0.6\pm 1.2)\times 10^3`$, where we have assumed the neutron polarization is $`\overline{P}=(96\pm 2)`$%. This is derived from our measurement of flipping ratio described in Section II A, with the assumption that the allowed range ($`93\%P100\%`$) spans $`2\sigma _{\overline{P}}`$.
Finally, we use the scaled results from the systematic test data (Figure 13) combined with Monte Carlo simulation studies to estimate the uncertainty of the Tilt-ATP systematic effect. For the test data, proton detector IV ($`\varphi _{\mathrm{IV}}=270^{}`$) was not operational. In calculating $`D`$ for the test data, only values from detectors I and III can therefore be used in Equation 21 with a result of $`\frac{1}{2}(D_\mathrm{I}+D_{\mathrm{III}})=(6.5\pm 1.4)\times 10^2`$. Monte Carlo simulations show that for a beam of radius 3 cm, the $`\mathrm{sin}(\varphi _\eta \varphi _\sigma )`$ behavior of Equation 18 is modified so that $`D_{test}=\frac{1}{2}(D_\mathrm{I}+D_{\mathrm{III}})/1.6=(4.1\pm 0.9)\times 10^2`$. This can be scaled by $`\mathrm{sin}\theta _\sigma `$, the ratio of polarization misalignments for the data and test runs. The individual values of $`D_\eta ^{l/s}`$ shown in Figure 14 are used to determine $`\theta _\sigma =(9\pm 3)\times 10^3`$ radians for the data run. This provides an upper limit for the uncertainty on the Tilt-ATP systematic effect of $`D`$(Tilt ATP)$`<D_{test}\mathrm{sin}\theta _\sigma 5.2\times 10^4`$. Though we use the test results to estimate this false $`D`$ effect, we expect the cancellation due to beam symmetry to be more complete for the data run because the test beam was intentionally distorted. We therefore consider this upper limit to be a conservative estimate of the largest possible false $`D`$ effect The contributions to the statistical and systematic uncertainties are given in Table II.
## V Summary and Conclusions
The apparatus used to perform a measurement of the $`D`$-coefficient in the beta-decay of polarized neutrons has been described. The data using the emiT detector have been analyzed using a technique that is insensitive to the nonuniform detection efficiency over the proton detectors. The initial run produced a statistically limited result of $`D=[0.6\pm 1.2(\mathrm{stat})\pm 0.5(\mathrm{syst})]\times 10^3`$. This result can be combined with earlier measurements to produce a new world average for the neutron $`D`$ coefficient of $`5.5\pm 9.5\times 10^4`$, which constrains the phase of $`g_A/g_V`$ to $`180.073^{}\pm 0.12^{}`$. This represents a 33% improvement (95% C.L.) over limits set by the current world average, and correspondingly further constrains standard model extensions with leptoquarks . The result is also interesting in light of upper limits provided by the neutron and <sup>199</sup>Hg electric dipole moments on T-odd, P-even interactions such as left-right symmetric models and exotic fermion models.
A second run is being planned with strategies to improve the statistical limitations related to background experienced in the first run. Our study of systematic effects presented here shows that the largest is the tilt-ATP effect. The uncertainty on this effect can be reduced significantly with more data taken in the transverse polarization mode described in Section IV C. With the planned improvements in place, it will be feasible to improve the sensitivity to $`D`$ to $`3\times 10^4`$ or less.
## Acknowledgments
We would like to thank Peter Herczeg for many helpful conversations. We are also grateful for the significant contributions of Steven Elliott. Thanks are due to Mel Anaya, Allen Myers, Tim Van Wechel, and Doug Will for technical support and to Vassilious Bezzerides, Laura Grout, Kyu Hwang, Christopher Scannell, Christina Scovel, and Kyle Sundqvist for their work on the project. We acknowledge the support of the National Institute of Standards and Technology, U.S. Department of Commerce, in providing the neutron facilities and other significant supplies and support. This research was made possible in part by grants from the U.S. Department of Energy Division of Nuclear Physics (contract numbers DE-AI05-93ER40784, DE-FG03-97ER41020, ยฟDE-AC03-76SF00098, and 00SCWE324) and the National Science Foundation. L.J.L. would also like to acknowledge support from the National Physical Science Consortium and the National Security Agency.
|
warning/0006/math0006180.html
|
ar5iv
|
text
|
# Infinitesimal aspects of the Laplace operator
## Introduction
Recall , , , that any manifold $`M`$, when seen in a model of Synthetic Differential Geometry (SDG), carries a reflexive symmetric relation $`_k`$ ($`k=0,1,2,\mathrm{}`$), where $`x_ky`$ reads โ$`x`$ and $`y`$ are $`k`$-neighboursโ; $`x_0y`$ means $`x=y`$; $`x_ky`$ implies $`x_{k+1}y`$. Also, $`x_ky,y_lz`$ implies $`x_{k+l}z`$. The set of $`(x,y)M\times M`$ with $`x_ky`$ is denoted $`M_{(k)}`$, the โ$`k`$โth neighbourhood of the diagonalโ, and for fixed $`x`$, the set $`\{yMy_kx\}`$ is denoted $`_k(x)`$ (โthe $`k`$-monad around $`x`$โ). In $`R^n`$, $`_k(0)`$ is denoted $`D_k(n)`$. Its elements $`u`$ are characterized by the condition that any homogeneous polynomial of degree $`k+1`$ vansihes on $`u`$. โ In this context, a Riemannian metric on $`M`$ can be given in terms of a map
$$g:M_{(2)}R$$
with $`g(x,x)=0`$, and with $`g(x,y)=g(y,x)`$ to be thought of as the โsquare-distance between $`x`$ and $`y`$โ, se , . (Also $`g`$ should be positive-definite, in a certain sense.)
Given a Riemannian metric $`g`$ on $`M`$, in this sense, one can construct the Levi-Civita connection , volume form , and hence also a notion of divergence of a vector field. And to a function $`f:MR`$, one can construct its gradient vector field, and hence one can construct the Laplacian $`\mathrm{\Delta }`$ by $`\mathrm{\Delta }(f)=div(grad(f))`$. This is what we shall not do here, rather, we shall exploit the richness of synthetic language to give a more economic and more geometric construction of $`\mathrm{\Delta }`$. The construction is more economic in the sense that the definition of $`\mathrm{\Delta }f(x)`$ only depends on knowing $`f`$ on a certain subset $`_L(x)_2(x)`$, where $`_2(x)`$ is what is required to make the usual $`divgrad`$ construction work, or for defining the individual terms in the formula $`\mathrm{\Delta }f(x)=^2f/x_i^2(x)`$.
The description of $`_L(x)M`$, or equivalently, the description of the $`L`$-neighbour relation $`_L`$, is coordinate free, see Definition 1 below, and therefore, too, is the description of $`\mathrm{\Delta }f`$ and of the notion of harmonic function. We get a characterization of harmonic functions, in terms of an average-value property, which is infinitesimal in character and does not involve integration, see Theorem 1 and Proposition 8.
In Section 3 we prove that diffeomorphisms which preserve the $`L`$-neighbour relation are precisely the conformal ones. Section 4 deals with the special case of the complex plane, and Section 5 explains the โsupport of the Laplacianโ in systematic algebraic terms.
## 1 Preliminaries
Although the notions we use are introduced in a coordinate free way, we have no intention of avoiding use of coordinates as a tool of proof. This Section contains in fact mainly certain coordinate calculations, which we believe will be useful also in other contexts where Riemannian geometry is treated in the present synthetic manner.
Working in coordinates in $`M`$ means that we are identifying (an open subset of) $`M`$ with (an open subset of) $`R^n`$; for simplicity, we talk about thse open subsets as if they were all of $`M`$ and $`R^n`$, respectively; all our considerations are anyway only local. The Riemannian metric $`g`$ on $`M`$ then becomes identified with a Riemannian metric on $`R^n`$, likewise denoted $`g`$, and it may be written (for $`x_2y`$) in the form of a matrix product,
$$g(x,y)=(yx)^TG(x)(yx),$$
where $`xyR^n`$ is viewed as a column matrix, and $`G(x)`$, for each $`x`$, is a symmetric positive definite $`n\times n`$ matrix.
Using coordinates, we may form affine combinations of (the coordinate sets of) points of $`M`$, at least for sufficiently nearby points, and such combinations will in general have only little geometric significance, since they depend on the choice of the coordinate system. However, we have the following useful fact:
###### Proposition 1
Assume $`y_1_1x`$ and $`y_2_1x`$ (so $`(x+y_2y_1)_2x`$). Then
$$g(x,x+y_2y_1)=g(y_1,y_2);$$
in particular, for $`x=0`$,
$$g(0,y_2y_1)=g(y_1,y_2).$$
Proof. We may assume $`x=0`$. Then $`g(0,y_2y_1)`$ and $`g(y_1,y_2)`$ are given, respectively, by
$$(y_2y_1)^TG(0)(y_2y_1)=2y_1^TG(0)y_2$$
and
$$(y_2y_1)^TG(y_1)(y_2y_1)=2y_1^TG(y_1)y_2.$$
Now expand $`G(y)`$ as $`G(0)+H(y)`$ where $`H`$ depends linearily on $`y_10`$. The difference between our two expressions is then $`2y_1^TH(y_1)y_2`$, which depends bilinearily on $`y_1`$ and therefore vanishes.
We shall see below (Proposition 3) that the โcoordinatewiseโ affine combination considered in Proposition 1 does have an invariant geometric meaning, provided the coordinate system is geodesic:
We say that the metric $`g`$ on $`R^n`$ (or equivalently, the coordinate system around $`x_0M`$) is geodesic at $`0R^n`$ (or at $`x_0M`$, respectively), if the first partial derivatives of $`G(x)`$, as functions of $`xR^n`$, vanish at $`0`$; equivalently, if $`G(x)=G(0)`$ for every $`x_10`$. (This is in turn equivalent to the vanishing at $`x_0`$ of the Christoffel symbols of the metric, in the given coordinate system.) It is classical that for every point $`x_0`$, there exists a coordinate system which is geodesic at $`x_0`$. If $`G(0)`$ is the identity matrix, one talks about a normal coordinate system at $`x`$, and such also exist. Cf. e.g. for such notions.
Recall from formula (2) that any Riemannian metric $`g:M_{(2)}R`$ admits a unique symmetric extension $`\overline{g}:M_{(3)}R`$; in coordinates it is given by
$$\overline{g}(x,y)=(yx)^T(G(x)+1/2(D_{(yx)}G)(x))(yx).$$
(1)
Recall also from Theorem 3.6 that for $`x_2z`$ in a Riemannian manifold, and for $`tR`$, there exists a unique $`y_0`$ with $`y_0_2x`$ and $`y_0_2z`$ which is a critical point for the function of $`y`$ given by
$$t\overline{g}(x,y)+(1t)\overline{g}(z,y);$$
(2)
We call this $`y_0`$ an (intrinsic) affine combination of $`x`$ and $`z`$. We write it $`tx+(1t)z`$; this raises a compatibility problem in case we are working in coordinates, since we can then also form the โalgebraicโ affine combination of two coordinate $`n`$-tuples. However, in geodesic coordinates at $`x`$, there is no problem, according to the following Proposition, which extends Proposition 3.7 in . Let us consider a coordinate system with $`x`$ identified with $`0`$.
###### Proposition 2
The critical point $`y_0`$ for the function in (2) is the algebraic affine combination $`tx+(1t)z`$, if either $`x_1z`$, or if the coordinate system is geodesic at $`x`$.
Proof. Since $`x`$ is identified with $`0`$ in the coordinate system, the affine combination in question is just $`(1t)z`$. To show that it is a critical value for (2) means that
$$t\overline{g}(0,(1t)z+v)+(1t)\overline{g}(z,(1t)z+v)$$
(3)
is independent of $`v_10`$. We write $`g`$ in terms of the symmetric matrices $`G`$, as above. Let us take a Taylor expansion of the function $`G(y)`$, writing
$$G(y)=G(0)+H(y),$$
where the entries of the matrix $`H(y)`$ are of degree $`1`$ in $`y`$; and if the coordinate system is geodesic at $`x=0`$, $`H(y)`$ is even of degree $`2`$ in $`y`$. We then calculate. We get a โsignificantโ part from each of the two terms in (3), and then some โerrorโ terms, each of which will vanish for degree reasons, as we shall argue.
The two significant terms are the two terms in
$$t(((1t)z+v)^TG(0)((1t)z+v))+(1t)((tz+v)^TG(0)(tz+v)).$$
Expanding out by bilinearity and symmetry, the terms involving $`v`$ linearly cancel each other; and the terms involving $`v`$ quadratically vanish because $`v_10`$. So the significant terms, jointly, do not depend on $`v_10`$.
The โerrorโ terms are of two kinds: partly, arising from the replacement of $`G(z)`$ by $`G(0)`$; here, $`H(z)`$ enters; and partly there are correction terms when passing from $`g`$ to $`\overline{g}`$ defined on pairs of third order neighbours. The error term of the first kind is a multiple of
$$(tz+v)^TH(z)(tz+v);$$
we expand this out by bilinearity, and use that $`H(z)`$ is of degree $`1`$, and $`v_10`$. We get four terms each of which vanish for degree reasons if either $`z_10`$ or if $`H(z)`$ is of degree $`2`$.
Finally, the correction terms for upgrading $`g`$ to $`\overline{g}`$ donโt occur if $`z_10`$, since then $`g`$ is only applied to pairs of second order neighbours. Thus the assertion of the Proposition about the case $`z_1x`$ is already proved. In general, the upgrading involves first partial derivatives of $`G`$, (see (1)), so in the case the coordinate system is geodesic at $`0`$, no correction term is needed for $`\overline{g}(0,(1t)z)`$, but only for $`\overline{g}(z,(1t)z+v)`$. Using the formula (1), we see that the correction needed is a certain multiple of
$$(tz+v)^T(D_{tz+v}G)(z)(tz+v),$$
hence a linear combination of terms
$$zD_zG(z)z,vD_zG(z)z,zD_vG(z)z,$$
and something that contains $`v`$ in a bilinear way. All these terms vanish for degree reasons: for, since $`H`$ vanishes in the first neighbourhood of $`0`$, $`D_vG(z)`$ is of degree $`1`$ in $`z`$, and $`D_zG(z)`$ is even of degree $`2`$ in $`z`$.
Essentially the same degree counting as in this proof gives the following result:
###### Lemma 1
Let $`y_1x`$ and $`z_2x`$; then using a geodesic coordinate system at $`x=0`$, the quantity $`\overline{g}(y,z)`$ may be calculated as $`(zy)^TG(0)(zy)`$.
Given a Riemannian manifold. If $`x_2z`$, the mirror image $`z^{}`$ of $`z`$ in $`x`$ is by definition the affine combination $`2xy`$, i.e. the $`y`$ which is critical value for $`2\overline{g}(x,y)\overline{g}(z,y)`$, Theorem 3.6. Also, the parallelogram formation $`\lambda `$ is descibed in . Finally, if $`t`$ is a tangent vector $`DM`$, its geodesic prolongation $`\overline{t}:D_2M`$ is determined by the validity, for all $`d_1,d_2D`$ of
$$\overline{t}(d_1+d_2)=\lambda (t(0),t(d_1),t(d_2)).$$
(Recall that $`DR`$ are the elements of square zero, $`D_2`$ the elements of cube zero.) Now Proposition 2 has the following Corollary:
###### Proposition 3
Let $`x_2z`$; then the mirror image $`z^{}`$ of $`y`$ w.r.to $`x`$ may be calculated as follows: take a geodesic coordinate system at $`x`$ with $`x=0`$. Then $`z^{}=z`$.
Let $`y_1x`$, $`z_1x`$. Then $`\lambda (x,y,z)`$ may be calculated as follows: take a geodesic coordinate system at $`x`$ with $`x=0`$. Then $`\lambda (x,y,z)=y+z`$.
Let $`t`$ be a tangent vector $`DM`$ at $`xM`$. Then the geodesic prolongation $`\overline{t}:D_2M`$ of $`t`$ may be calculated as follows: take a geodesic coordinate system at $`x`$ with $`x=0`$. Let $`u`$ be the unique vector in $`R^n`$ so that $`t(d)=du`$ for all $`dD`$. Then for $`\delta D_2`$, $`\overline{t}(\delta )=\delta u`$
(The vector $`uR^n`$ appearing in the last clause is usually called the principal part of $`t`$, relative to the coordinate system.)
If $`t`$ and $`s`$ are tangent vectors at the same point $`x`$ of a Riemannian manifold $`M,g`$, we define their inner product $`<t,s>`$ by the validity, for all $`d_1d_2D`$, of
$$d_1d_2<t,s>=\frac{1}{2}g(t(d_1),s(d_2)).$$
(4)
In this way, the tangent vector space $`T_xM`$ is made into an inner product space, (and this is the contact point with the classical formulation of Riemannian metric).
If $`u`$ and $`v`$ are the principal parts of tangent vectors $`t`$ and $`s`$ at $`xM`$, in some coordinate system at $`x=0`$ (not necessarily geodesic), one has
$`<t,s>=u^TG(0)v`$; this follows easily from Proposition 1.
Combining Lemma 1 and Proposition 3, one gets
###### Lemma 2
Let $`t`$ be a tangent vector. Then for $`dD`$, $`\delta D_2`$, we have
$$\overline{g}(t(d),\overline{t}(\delta ))=(\delta ^22d\delta )<t,t>.$$
We are going to define the orthogonal projection of $`z`$ ($`z_2x`$) onto a proper tangent $`t`$ at $`x`$. We first define the scalar component of $`z`$ along $`t`$; this is unique number $`\alpha (z,t)`$ so that
$$d\alpha (z,t)=\frac{1}{2}\frac{(g(x,z)\overline{g}(t(d),z))}{<t,t>}$$
(5)
for all $`dD`$. Note that if $`z=x`$, $`\alpha (z,t)=0`$, and from this follows that for any $`z_2x`$, $`\alpha (z,t)_20`$, in other words $`\alpha (z,t)D_2`$. From Lemma 2, applied twice (once with $`d=0`$, once with a general $`dD`$), it is immediate to deduce that if $`z`$ is of the form $`\overline{t}(\delta )`$ for a $`\delta D_2`$, then $`\alpha (z,t)=\delta `$.
We define the orthogonal projection $`proj_t(z)`$ by
$$proj_t(z)=\overline{t}(\alpha (z,t)).$$
Note that it is a second-order neighbour of $`x`$. It follows from the above that if $`z`$ is of the form $`\overline{t}(\delta )`$, then $`proj_t(z)=z`$.
## 2 Laplacian neighbours
Here is the crucial definition:
###### Definition 1
Let $`z_2x`$. We say that $`z`$ is a Laplacian neighbour of $`x`$ (written $`z_Lx`$) if for every proper tangent $`t`$ at $`x`$, we have
$$g(x,z)=ng(x,proj_t(z)),$$
(6)
where $`n`$ is the dimension of the manifold.
Maybe one of the names โisotropic, harmonic, or conformal, neighbourโ would be more appropriate.
Clearly $`z_1x`$ implies $`z_Lx`$; for if $`z`$ is a first-order neighbour of $`x`$, then so is its orthogonal projection, and hence both the $`g`$-quantities to be compared in (6) are zero. If the dimension $`n`$ is 1, $`_L`$ is the same as $`_2`$ but in general, the set $`_L(x)`$ of $`L`$-neighbours of $`x`$ is much smaller than the set $`_2(x)`$ of second-order neighbours; in fact, the ring of functions on $`_L(x)`$ is a finite dimensional vector space which is just one dimension bigger than the ring of functions on $`_1(x)`$, as we shall see in the proof of Proposition 5 below.
We conjecture that the relation $`_L`$ is symmetric, but we havenโt been able to do the necessary calculations, except in the case of $`R^n`$, where the symmetry is easy to prove, using Proposition 4 below.
Note the following curious phenomenon in dimension $`n2`$: if $`z_Lx`$, then $`z`$ does not connect to $`x`$ by any geodesic $`D_2M`$ (given by a proper tangent vector $`t`$), except perhaps in the trivial case when $`g(x,z)=0`$. In other words, the $`L`$-neighbours of $`x`$ are genuinely isotropic, in the sense that they are in no preferred direction $`t`$ (hence the alternative name โisotropic neighbourโ suggested). Nevertheless, there are sufficiently many $`L`$-neighbours of $`x`$ to define the Laplacian differential operator $`\mathrm{\Delta }`$, see Theorem 1 below.
Let us assume the manifold in question has dimension $`n`$. Then we have
###### Proposition 4
In any geodesic normal coordinate system at $`x=0`$, $`z=(z_1,\mathrm{},z_n)`$ is $`_L0`$ if and only if
$$z_i^2=z_j^2\text{ for all }i,j\text{, and }z_iz_j=0\text{ for }ij$$
(and $`z_i^3=0`$ for all $`i`$; this latter condition follows from the other two if $`n2`$).
Proof. First, if $`t`$ and $`s`$ are tangent vectors at $`x=0`$ with principal parts $`u`$ and $`v`$, respectively (meaning $`t(d)=du,s(d)=dv`$), then $`<t,s>=uv`$, where $``$ denotes the usual dot product of vectors in $`R^n`$. Also, if $`t`$ is a tangent at $`x=0`$ with principal part $`u`$, then $`\alpha (z,t)=(zu)/(uu)`$; for, calculating the enumerator in (5) gives (using Proposition 1)
$$zz\overline{g}(du,z)=zz(zdu)(zdu)=2dzu.$$
From the third clause in Proposition 3, we then get the familiar looking
$$proj_t(z)=\frac{zu}{uu}u.$$
(7)
In particular, if $`t`$ is the (proper) tangent vector with principal part $`e_iR^n`$ ($`=(0,\mathrm{},1,\mathrm{}0)`$ (with 1 in the $`i`$โth position, $`0`$โs elsewhere), then $`proj_t(z_1,\mathrm{},z_n)=z_ie_i`$. In particular $`g(0,proj_t(z))=z_i^2`$. If, on the other hand, $`t`$ is the tangent vector with principal part $`e_{i,j}`$ (the vector with $`1`$โs in the $`i`$โth and in the $`j`$โth position, $`ij`$, $`0`$โs elsewhere), then $`proj_t(z)`$ has $`(z_i+z_j)/2`$ in the $`i`$โth and in the $`j`$โth position, and $`0`$โs elsewhere. In particular,
$$g(0,proj_t(z))=\frac{1}{2}(z_i^2+z_j^2)+z_iz_j.$$
If $`z`$ therefore is an $`L`$-neighbour of $`0`$, we conclude that $`z_i^2=z_j^2`$ for all $`i,j`$, and that $`z_iz_j=0`$ if $`ij`$.
Conversely, assume that in some geodesic normal coordinate system at $`x=0`$, the coordinates $`(z_1,\mathrm{},z_n)`$ satisfy the equations $`z_i^2=z_j^2`$, $`z_iz_j=0`$ for $`ij`$, and let $`t`$ be a proper tangent vector at $`x`$ with principal part $`u=(u_1,\mathrm{},u_n)`$. Then
$$proj_t(z)=\frac{zu}{uu}u,$$
and therefore
$$g(x,proj_t(z))=(\frac{zu}{uu}u)(\frac{zu}{uu}u),$$
which we calculate by arithmetic to be
$$\frac{(_iu_iz_i)(_ju_jz_j)}{uu}=\frac{_{ij}u_iu_jz_iz_j}{uu},$$
but since $`z_iz_j=0`$ for $`ij`$, only the โdiagonalโ terms survive, and we are left with
$$\frac{_iu_iu_iz_iz_i}{uu}.$$
But $`z_iz_i=z_1z_1`$ for all $`i`$, so this factor can go outside the sum sign in the enumerator, and we get $`z_1^2(_iu_iu_i)/uu=z_1^2`$, which is $`1/n`$ times $`z_i^2`$ since all the $`z_i^2`$ are equal. This proves the Proposition.
From Propositions 3 and 4, one immediately deduces that if $`z_Lx`$, then also $`z^{}_Lx`$ for any affine combination $`z^{}=tx+(1t)z`$ ($`tR`$).
###### Proposition 5
If two functions $`f_1`$ and $`f_2`$: $`_L(x)R`$ agree on $`_1(x)`$ there is a unique number $`cR`$ such that for all $`z_Lx`$
$$f_1(z)f_2(z)=cg(x,z).$$
Proof. Using a geodesic normal coordinate system at $`x=0`$, it is a matter of analyzing the ring of functions $`_L(0)R`$ for the case where $`M=R^n`$ with standard inner-product metric. The Proposition gives that $`_L(0)`$ may be described as $`D_L(n)R^n`$, defined by
$$D_L(n):=\{(d_1,\mathrm{},d_n)R^nd_i^2=d_j^2\text{, and }d_id_j=0\text{ for }ij\},$$
(8)
(for $`n2`$; for $`n=1`$, $`D_L(1)=D_2=\{\delta R\delta ^3=0\}`$). This is (for $`n2`$) the object represented by the Weil algebra $`๐ช(D_L(n)):=k[Z_1,\mathrm{},Z_n]/I`$, where $`I`$ is the ideal generated by the $`Z_i^2Z_j^2`$, and by $`Z_iZ_j`$ for $`ij`$. It is immediate to calculate that, as a vector space, this ring is $`(n+2)`$-dimensional, with linear generators
$$1,Z_1,\mathrm{},Z_n,Z_1^2+\mathrm{}+Z_n^2.$$
By the general (Kock-Lawvere) axiom scheme for SDG , , this means that any function $`f:D_L(n)R`$ is of the form
$$f(z_1,\mathrm{},z_n)=a+\underset{i}{}b_iz_i+c(\underset{i}{}z_i^2),$$
for unique $`a,b_1,\mathrm{},b_n,cR`$, or equivalently
$$f(z_1,\mathrm{},z_n)=a+\underset{i}{}b_iz_i+cg(0,z).$$
Since the restriction of $`f`$ to $`D(n)`$ is given by the data $`a,b_1,\mathrm{},b_n`$, the unique existence of $`c`$ follows.
The following Theorem deals with an arbitrary Riemannian manifold $`M,g`$ of dimension $`n`$, and gives a coordinate free characterization of the Laplacian operator $`\mathrm{\Delta }`$.
###### Theorem 1
For any $`f:_L(x)R`$, there is a unique number $`L`$ with the property that for any $`z_Lx`$
$$f(z)+f(z^{})2f(x)=Lg(x,z),$$
where $`z^{}`$ denotes the mirror image of $`z`$ in $`x`$. We write $`\mathrm{\Delta }f(x):=nL`$.
Put differently,
$$f(z)+f(z^{})2f(x)=\frac{\mathrm{\Delta }f(x)}{n}g(x,z).$$
If the function $`f`$ is harmonic at $`x`$, meaning that $`\mathrm{\Delta }f(x)=0`$, it follows that it has a strong average value property: the value at $`x`$ equals the average value of $`f`$ over any pair of points $`z`$ and $`z^{}`$ ($`L`$-neighbours of $`x`$) which are symmetrically located around $`x`$.
Proof. Again, we pick a normal geodesic coordinate system at $`x=0`$, so identify $`_L(x)`$ with $`D_L(n)`$; then $`z^{}`$ gets identified with $`z`$, by Proposition 3. The left hand side of the expression in the Theorem then has restriction 0 to $`D(n)`$, being (with notation as above) $`(a+b_iz_i)+(a+b_i(z_i))2a.`$ Hence the unique existence of $`L`$ follows from Proposition 5.
The following Proposition serves to as partial justification of the use of the name โLaplacianโ for the $`\mathrm{\Delta }`$ considered in the Theorem. We consider the standard Riemannian metric on $`R^n`$, $`g(x,z)=zx^2`$, for $`z_2x`$.
###### Proposition 6
Let $`f:R^nR`$ and let $`xR^n`$. Then
$$\mathrm{\Delta }f(x)=\underset{i}{}\frac{^2f}{x_i^2}(x).$$
Proof. For simplicity, let $`x=0`$, so that $`z^{}=z`$. We Taylor expand $`f(z)`$ and $`f(z)`$ from $`0`$, and consider $`f(z)+f(z)2f(0)`$; then terms of degree $`1`$ cancel, and we get
$$\underset{ij}{}\frac{^2f}{x_ix_j}z_iz_j+\text{ higher terms};$$
now the calculation proceeds much like the one in the proof of Proposition 4 above: if $`z_L0`$, only the diagonal terms in the sum survive, all the $`z_i^2`$ are equal to $`z_1^2`$, which we move outside the parenthesis, and get $`z_1^2`$ times the classical Laplacian $`^2f/x_i^2(0)`$. But $`z_1^2=1/ng(0,z)`$.
Similarly, one proves by Taylor expansion
###### Proposition 7
If $`z_Lx`$ in $`R^n`$, then for any $`f:R^nR`$,
$$f(z)=f(x)+df_x(zx)+\frac{1}{2n}\mathrm{\Delta }f(x)zx^2.$$
Recall that any function $`MR`$ looks affine on any 1-monad $`_1(x)`$; functions that look affine on the larger $`L`$-monads $`_L(x)`$ are precisely the harmonic ones:
###### Proposition 8
Assume $`f:MR`$ is harmonic at $`x`$. Then for an $`z_Lx`$, $`f`$ preserves affine combinations of $`x`$ and $`z`$. Conversely, if for given $`x`$, $`f`$ preserves affine combinations of $`x`$ and $`z`$ for every $`z_Lx`$, then $`f`$ is harmonic at $`x`$. (โHarmonicโ at $`x`$ here in the sense: $`\mathrm{\Delta }f(x)=0`$.)
(Recall that the affine combination $`tx+(1t)z`$ is defined as the critical point $`y_0`$ in (2).)
Proof. Assume $`z_Lx`$, and pick a geodesic normal coordinate system at $`x=0`$. Without loss of generality, we may assume $`f(0)=0`$. Then to say that $`f`$ preserves affine combinations of $`x`$ and $`z`$ is to say that for all $`sR`$, $`f(sz)=sf(z)`$. For $`z_L0`$, we have by Proposition 7 that $`f(z)=a_iz_i+cz_i^2`$ for unique $`a_i`$ and $`c`$ ($`c=\mathrm{\Delta }f(0)/2n`$); $`f(sz)`$ and $`sf(z)`$ have the same terms of first order in $`z`$; their second order terms are respectively $`cs^2z_i^2`$ and $`scz_i^2`$, and if these two expressions are to be equal for all $`s`$ and all $`z`$, we must have $`c=0`$. This means that $`f`$ is harmonic at $`x`$. Conversely, if $`f`$ preserves affine combinations of the kind mentioned, it preserves the affine combination $`2xz`$, or equivalently, the left hand side of the expression in Theorem 1 is 0, hence it follows that $`L`$ and hence $`\mathrm{\Delta }f(x)`$ is 0.
Remark. With some hesitation, I propose to call a map $`f:MN`$ between Riemannian manifolds harmonic if it preserves affine combinations of $`L`$-neighbours in $`M`$ and if it preserves the property of being $`L`$-neighbours. I have not been able to compare the proposed definition, with a certain classical concept of harmonic map between Riemannian manifolds. But at least: When the codomain is $`R`$, the definition is the basic classical one of harmonic function, by Proposition 8. For, preservation of the $`_L`$ relation is automatic when the codomain is $`R`$, since in $`R`$, $`_L`$ is the same as $`_2`$.
## 3 Conformal maps
We consider a diffeomorphism $`f:MN`$ between Riemannian manifolds $`(M,g),(N,h)`$. To say that $`f`$ is an isometry at $`xM`$ is to say that for all $`z_2x`$, $`g(x,z)=h(f(x),f(z))`$. To say that $`f`$ is conformal at $`xM`$ with constant $`k=k(x)>0`$ is to say that for all $`z_2x`$, $`h(f(x),f(z))=k(x)g(x,z)`$ (so if $`k(x)=1`$, $`f`$ is an isometry at $`x`$). The terminology agrees with classical usage, as we shall see below. We first prove
###### Proposition 9
Assume $`f:MN`$ is conformal at $`xM`$ with
$`h(f(x),f(z))=kg(x,z)`$. Then for all $`y_1_1x`$, $`y_2_1x`$,
$$h(f(y_1),f(y_2))=kg(y_1,y_2),$$
and conversely.
Proof. We choose coordinates, and assume $`x=0`$ and $`f(x)=0`$; the metrics in $`M`$ and $`N`$ are then given by functions $`g`$ and $`h`$, respectively, and they are in turn given by symmetric matrices $`G(y)`$, and $`H(z)`$ for all $`yM`$ and $`zN`$. We now calculate $`kg(y_1,y_2)`$. We have, by Proposition 1 that
$$kg(y_1,y_2)=kg(0,y_2y_1)=h(0,f(y_2y_1)).$$
Now there is a bilinear $`B(,)`$ such that for all pairs of $`1`$-neighbours $`y_1,y_2`$ of $`0`$, we have $`f(y_2y_1)=f(y_2)f(y_1)+B(y_1,y_2)`$. So the calculation continues
$$=h(0,f(y_1)f(y_2)+B(y_1,y_2))$$
$$=(f(y_1)f(y_2)+B(y_1,y_2))^TH(0)(f(y_1)f(y_2)+B(y_1,y_2)).$$
Since $`f`$ depends in a linear way of $`y_1_10`$ and $`y_2_10`$, this whole expression multiplies out by linearity, and for degree reasons all terms involving $`B`$, as well as some others, vanish, and we are left with $`2f(y_1)^TH(0)f(y_2)`$. On the other hand, $`h(f(y_1),h(f(y_2))=h(0,f(y_2)f(y_1))`$, by Proposition 1, and writing this in terms of $`H(0)`$ gives the same expression.
The converse is proved in the same way for $`z_20`$ of the form $`y_2y_1`$ with $`y_1_10`$ and $`y_2_10`$, but this suffices to get the result for all $`z_20`$, by general principles of SDG (โ$`R`$, and hence any manifold, perceives the addition map $`D(n)\times D(n)D_2(n)`$ to be epicโ.)
Call a linear map $`F:UV`$ between inner product vector spaces conformal with constant $`k>0`$ if for all $`u_1,u_2U`$
$$<F(u_1),F(u_2)>=k<u_1,u_2>.$$
It follows immediately from Proposition 9, and from the construction of inner product in the vector space of tangents at $`x`$, and at $`f(x)`$, that if $`f`$ is conformal at $`x`$ with constant $`k`$, then $`df_x:T_xMT_{f(x)}N`$ is a conformal linear map with the same constant $`k`$. The converse also holds; for if $`df_x`$ is conformal with constant $`k`$, we deduce that for all pairs of tangents $`t`$ and $`s`$ at $`x`$
$$g(f(t(d_1)),f(s(d_2)))=kg(t(d_1),s(d_2)),$$
and hence
$$g(f(y_1)),f(y_2))=kg(y_1,y_2)$$
(9)
for all $`y_i`$โs of the form $`t(d)`$ for a tangent vector $`t`$ and a $`dD`$. Again by general principles, any manifold $`N`$ โperceives all 1-neighbours of $`x`$ to be of this formโ. From Proposition 9 we therefore deduce that $`f`$ is conformal at $`x`$ with constant $`k`$.
###### Theorem 2
A diffeomorphism $`f`$ is conformal at $`xM`$ if and only if $`f`$ maps $`_L(x)`$ into $`_L(f(x))`$.
Proof. Assume $`f`$ maps $`_L(x)`$ into $`_L(f(x))`$. We may pick normal coordinates at $`x`$ as well as at $`f(x)`$. The neighbourhoods $`_L(x)`$ and $`_L(f(x))`$ then both get identified with $`D_L(n)`$, and $`x=0`$, $`f(x)=0`$. The restriction of $`f`$ to $`D_2(n)`$, $`f:D_2(n)R^n`$, takes $`0`$ to $`0`$ and is therefore of the form $`f(y)=Ay+B(y)`$, where $`A`$ is an $`n\times n`$ matrix, and $`B(y)`$ is a map $`R^nR^n`$ which is homogeneous of degree 2 in $`yR^n`$, i.e. an $`n`$-tuple of quadratic forms $`B_i`$.
Assume now that $`f`$ maps $`D_L(n)`$ into itself. For $`zD_L(n)`$, the $`i`$โth coordinate of $`f(z)`$ is
$$f_i(z)=\underset{k}{}a_{ik}z_k+B_i(z).$$
Squaring this, only the terms in $`(_ka_{ik}z_k)(_la_{il}z_l)`$ survive for degree reasons (using that $`zD_2(n)`$). But using further that $`z_kz_l=0`$ for $`kl`$, only the โdiagonalโ terms survive, and we get
$$f_i(z)^2=\underset{k}{}a_{ik}^2z_k^2=z_1^2\underset{k}{}a_{ik}^2.$$
(10)
Similarly for $`ij`$
$$f_i(z)f_j(z)=z_1^2(\underset{k}{}a_{ik}a_{jk}).$$
(11)
If now $`f(z)D_L(n)`$ for all $`zD_L(n)`$, we get that the expression in (10) is independent of $`i`$, and from the uniqueness assertion in Proposition 5 we therefore conclude
$$\underset{k}{}a_{ik}^2=\underset{k}{}a_{jk}^2\text{ for all }i,j;$$
and similarly we conclude from (11) that
$$\underset{k}{}a_{ik}a_{jk}=0\text{ for }ij.$$
These two equations express that all the rows of the matrix $`A`$ have the same square norm $`k`$, and that they are mutually orthogonal. This implies that the linear map $`df_x`$ represented by the matrix is conformal, and hence $`f`$ is conformal at $`x`$.
The proof that conformality of $`f`$ at $`x`$ implies that $`f`$ maps $`_L(x)`$ into $`_L(f(x))`$ goes essentially through the same calculation, and is omitted.
## 4 A famous pseudogroup in dimension 2
The content of the present section is partly classical, namely the equivalence of the various ways of describing the notion of holomorphic map from (a region in) the complex plane $`C=R^2`$ to itself. Synthetic concepts enter essentially in two of the conditions in the Theorem below, namely 1) and 7).
An almost complex structure on a general manifold $`M`$ consists in giving, for each $`xM`$, a map $`I_x:_1(x)_1(x)`$ with $`I_x(x)=x`$ and $`I_x(I_x(z))=z^{}`$ for any $`z_1x`$; Here, $`z^{}`$ denotes the mirror image of $`z`$ in $`x`$, i.e. the affine combination $`2xz`$; recall that affine combinations of 1-neighbours make โabsolutelyโ sense, i.e. do not depend on, say, a Riemannian structure. It is clear what it means for a map $`f`$ to preserve such structure at the point $`x`$: $`f(I_x(z))=I_{f(x)}(f(z))`$.
The manifold $`R^2`$ carries a canonical almost-complex structure, given by
$$I_{(x_1,x_2)}(z_1,z_2)=(x_1(z_2x_2),x_2+(z_1x_1)).$$
Identifying $`R^2`$ with the complex plane $`C`$, this is just
$$I_x(z)=x+i(zx).$$
Utilizing the multiplication of the complex plane $`C`$, we may consider the set $`D_C`$ of elements of square zero in $`C`$ (recalling the fundamental role which the set $`D`$ of elements of square zero in $`R`$ plays in SDG). We have, by trivial calculation,
###### Proposition 10
Under the identification of $`C`$ with $`R^2`$,
$$D_C=D_L(2).$$
Having $`D_C`$, we may mimick the basics of SDG and declare a function $`f:CC`$ to be complex differentiable at $`xC`$ if there is a number $`f^{}(x)C`$ so that
$$f(z)=f(x)+f^{}(x)(zx)\text{ for all }z\text{ with }zxD_C.$$
(The uniqueness of such $`f^{}(x)`$, justifying the notation, follows from the general axiom scheme of SDG, applied to $`D_1(2)`$, the 1-jet classifier in $`R^2`$. Note $`D_1(2)D_L(2)`$.) โ The notion of course makes sense for functions $`f`$ which are just defined locally around $`x`$.
###### Theorem 3
Let $`f:R^2R^2`$ be a (local) orientation preserving diffeomorphism. Let $`xR^2=C`$. Then the following conditions are equivalent:
1) $`f`$ maps $`_L(x)`$ into $`_L(f(x))`$
2) $`f`$ is conformal at $`x`$
3) $`f`$ satisfies Cauchy-Riemann equations at $`x`$
4) $`f`$ preserves almost complex structure at $`x`$.
Also the following conditions are equivalent, and they imply 1)-4):
5) $`f`$ is complex-differentiable at $`x`$
6) $`f`$ maps $`_L(x)`$ into $`_L(f(x))`$, and $`f`$ preserves affine combinations of $`x`$ and $`z`$ for any $`z_Lx`$.
Finally, if 1)-4) hold for all $`x`$, 5) and 6) hold for all $`x`$.
(Note that 6) says that $`f`$ is harmonic at $`x`$, in the sense of Remark at the end of Section 2.)
Proof. The equivalence of 1) and 2) is already in Theorem 2, and this in turn is, as we have seen, equivalent to conformality of the linear $`df_x`$. But conformal orientation preserving $`2\times 2`$ matrices are of the form
$$\left[\begin{array}{cc}\hfill a& \hfill b\\ \hfill b& \hfill a\end{array}\right].$$
(12)
Since the entries of the matrix for $`df_x`$ are $`f_i/x_j`$, this form (12) of the matrix therefore expresses that the Cauchy-Riemann equations hold at $`x`$, i.e. is equivalent to 3). On the other hand, a simple calculation with $`2\times 2`$ matrices give that a matrix commutes with the matrix $`I=\left[\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right]`$ for the almost complex structure iff it has the above โCauchy-Riemannโ form (12).
Now assume 5). If $`f`$ is complex differentiable at $`x`$, we prove that condition 1) holds at $`x`$ as follows. Let $`z_Lx`$. Then $`(zx)^2=0`$ Proposition 10, and by complex differentiability
$$f(z)f(x)=f^{}(x)(zx),$$
(13)
so since the right hand side has square zero, then so does the left hand side, but again by Proposition 10, this means that $`f(z)_Lf(x)`$, proving 1), and hence also the first part of 6). But also, if $`f`$ is complex-differentiable at $`x`$, $`f`$ preserves affine combinations of the form $`tx+(1t)z`$ for $`z_L(x)`$; this follows from (13), since $`zxD_C(x)`$ by Proposition 10, so also the second part of 6) is proved. Conversely, if 6) holds, $`f`$ is conformal at $`x`$ by Theorem 2, so $`df_x`$ is if the form (12). Then $`f^{}(x)=a+ib`$ will serve as the complex derivative; for, since $`f`$ preserves affine combinations of $`x`$ and $`z`$ for $`z_Lx`$, we have the first equality sign in
$$f(z)=f(x)+df_x(zx)=f(x)+f^{}(x)(zx)$$
for such $`z`$, i.e. for $`zxD_L(2)=D_C`$.
Finally, assume 1)-4) hold for all $`x`$. Then we may differentiate the Cauchy-Riemann equations for $`f=(f_1,f_2)`$ by $`/x_1`$ and $`/x_2`$ and compare, arriving in the standard way to $`\mathrm{\Delta }f_10`$ and $`\mathrm{\Delta }f_20`$. From the โTaylor expansionโ in Proposition 7, applied to $`f_1`$, we conclude that $`f_1(z)=f_1(x)+(df_1)_x(zx)`$ for $`z_Lx`$, and similarly for $`f_2`$, so $`f(z)=f(x)+df_x(zx)`$ for such $`z`$, i.e. for $`zxD_C`$. Since $`df_x`$ is given by a conformal matrix (12), by 2), this proves that $`a+ib`$ will serve as the complex derivative of $`f`$ at $`x`$.
## 5 Support of the Laplacian
We arrived at $`D_L(n)`$ from the geometric side, namely as the $`_L`$-neighbours of $`0`$ in the Riemannian manifold $`M=R^n`$; the differential operator $`\mathrm{\Delta }`$ was then seen to provide the top term in the Taylor expansion of functions defined on $`D_L(n)`$.
Here, we briefly indicate how to arrive at $`D_L(n)`$ from the algebraic side, starting with $`\mathrm{\Delta }=^2/x_i^2`$. More precisely, we consider $`\mathrm{\Delta }`$ as a distribution at $`0R^n`$. So $`\mathrm{\Delta }`$ is the linear map
$$k[X_1,\mathrm{},X_n]R$$
given by
$$f\underset{i}{}\frac{^2f}{x_i^2}(0).$$
(14)
The algebraic concept that will give $`D_L(n)`$ out of this data is the notion of coalgebra, and subcoalgebra, as in . If we let $`A`$ denote the algebra $`k[X_1,\mathrm{},X_n]`$, then the distribution $`\mathrm{\Delta }`$ of (14) factors
$$ABk,$$
where $`AB`$ is an algebra map, and $`B`$ is finite dimensional (take e.g. $`B=A/J`$ where $`J`$ is the ideal generated by monomials of degree $`3`$). The set $`A^o`$ of linear maps $`Ak`$ having such a factorization property constitute a coalgebra, Proposition 6.0.2. Every element in a coalgebra generates a finite dimensional subcoalgebra, by Theorem 2.2.1. In particular, $`\mathrm{\Delta }A^o`$ generates a finite dimensional coalgebra $`[\mathrm{\Delta }]`$ of $`A^o`$, and this coalgebra โisโ $`D_L(n)`$. More specifically, the dual algebra of $`[\mathrm{\Delta }]`$ is the coordinate ring $`๐ช(D_L(n))`$ of $`D_L(n)`$, i.e. the Weil algebra $`๐ช(D_L(n)):=k[Z_1,\mathrm{},Z_n]/I`$ considered in the proof of Proposition 5, as we shall now argue.
The following โLeibniz ruleโ for $`\mathrm{\Delta }`$ is well known,
$$\mathrm{\Delta }(fg)=\mathrm{\Delta }fg+2\underset{i}{}\frac{f}{x_i}\frac{g}{x_i}+f\mathrm{\Delta }g.$$
This means that in the coalgebra $`A^o`$, we have the following formula for $`\psi (\mathrm{\Delta })`$ ($`\psi =`$ the comultiplication of the coalgebra; $`\delta `$ the Dirac distribution โevaluate at 0โ):
$$\psi (\mathrm{\Delta })=\mathrm{\Delta }\delta +2\underset{i}{}\frac{}{x_i}\frac{}{x_i}+\delta \mathrm{\Delta },$$
(15)
where now $`\mathrm{\Delta }`$, $`d/x_i`$, $`\delta `$ are viewed as distributions at $`0`$, like in (14), meaning that one evaluates in $`0`$ after application,
$$f(\mathrm{\Delta }f)(0)\text{}f\frac{f}{x_i}(0)\text{}ff(0).$$
From (15), (and from $`\psi (/x_i)=/x_i\delta +\delta /x_i`$, which expresses the Leibniz rule for $`/x_i`$) we see that the subcoalgebra $`[\mathrm{\Delta }]`$ generated by $`\mathrm{\Delta }`$ is generated as a vector space by the elements
$$\delta \text{}\frac{}{x_1},\mathrm{},\frac{}{x_n}\text{}\mathrm{\Delta },$$
and since these are clearly linearly independent, we see that $`[\mathrm{\Delta }]A^o`$ is $`(n+2)`$-dimensional. The dual algebra of $`[\mathrm{\Delta }]`$ is a quotient algebra $`A/I`$ of $`A`$, where $`I`$ is the ideal of those $`fA`$ which are annihilated by the elements of $`[\mathrm{\Delta }]`$. This ideal $`I`$ contains $`x_i^2x_j^2`$, and $`x_ix_j`$ for $`ij`$. Since the quotient of $`A`$ by the ideal generated by $`x_i^2x_j^2`$, and $`x_ix_j`$ for $`ij`$ is already $`(n+2)`$-dimensional, as calculated in the proof of Proposition 5, it follows that the quotient algebra there is actually the dual of $`[\mathrm{\Delta }]`$.
The idea that a coalgebra like $`[\mathrm{\Delta }]`$ is itself an infinitesimal geometric object goes back to Gavin Wraith in the early seventies, . The specific way of generating Weil algebras from differential operators was considered by Emsalem (without coalgebras).
|
warning/0006/gr-qc0006083.html
|
ar5iv
|
text
|
# Symmetries of Bianchi I space-times
## 1 Introduction
Collineations are geometrical symmetries which are defined by the general relation:
$$_\xi ๐ฝ=๐ฒ$$
(1)
where $`๐ฝ`$ is any of the quantities $`g_{ab},\mathrm{\Gamma }_{bc}^a,R_{ab},R_{bcd}^a`$ and geometric objects constructed by them and $`๐ฒ`$ is a tensor with the same index symmetries as $`๐ฝ`$. Some of the well known (and most important) types of collineations are: Conformal Killing vector (CKV) $`\xi ^a`$ defined by the requirement $`_\xi g_{ab}=2\psi g_{ab}`$ and reducing to a Killing vector (KV) when $`\psi =0`$, to a Homothetic vector field (HVF) when $`\psi =`$const. and to a Special Conformal Killing vector (SCKV) when $`\psi _{;ab}=0`$. Proper Affine Conformal vector (ACV) is defined by the requirement $`_\xi g_{ab}=2\psi g_{ab}+2H_{ab}`$ where $`H_{ab;c}=0,`$ $`\psi _{;ab}0`$ and reducing to an Affine Vector (AV) when $`\psi _{,a}=0`$ and to a Special Affine Conformal vector (SACV) when $`\psi _{;ab}=0`$. Curvature Collineation (CC) defined by the requirement $`_\xi R_{bcd}^a=0`$ and finally Ricci Collineation (RC) defined by the requirement $`_\xi R_{ab}=0`$.
Collineations other than motions (KVs) can be considered as non-Noetherian symmetries and can be associated with constants of motion and, up to the level of CKVs, they can be used to simplify the metric . For example AVs are related to conserved quantities (a result used to integrate the geodesics in FRW space-times), RCs are related to the conservation of particle number in FRW space-times and the existence of CCs implies conservation laws for null electromagnetic fields .
The set of (smooth) collineations of a space-time can be related with an inclusion relation leading to a tree like inclusion diagram which shows their relative hierarchy. A collineation of a given type is proper if it does not belong to any of the subtypes in this diagram. In order to relate a collineation to a particular conservation law and its associated constant(s) of motion the properness of the collineation must first be assured.
A different type of symmetry we shall discuss, which is of a kinematic nature, is the Kinematic Self Similarity (KSS). It is defined by the requirements $`_\xi u_a=\alpha u_a`$, $`_\xi h_{ab}=2\delta h_{ab}`$ where $`u^a`$ is the four velocity of the fluid, $`h_{ab}=g_{ab}+u_au_b`$ $`(u_au^a=1)`$ is the projection tensor normal to $`u^a`$ and $`\alpha ,\delta `$ are constants. A KSS reduces to a HVF when $`\alpha =\delta 0`$ and to a KV when $`\alpha =\delta =0.`$ The KSS are characterised by the scale independent ratio $`\alpha /\delta `$, which is known as the similarity index. When $`\alpha =0`$ the KSS is of type zero (zeroth kind) and when $`\delta =0`$ it is of type infinite. Kinematic Self Similarity should be regarded as the relativistic generalisation of self similarity of Newtonian Physics rather than the generalisation of the space-time homotheties. From the physical point of view the detailed study of cosmological models admitting KSS shows that they can represent asymptotic states of more general models or, under certain conditions, they are asymptotic to an exact homothetic solution .
A diagonal Bianchi I space-time is a spatially homogeneous space-time which admits an abelian group of isometries $`G_3`$, acting on spacelike hypersurfaces, generated by the spacelike KVs $`\xi _1=_x,\xi _2=_y,\xi _3=_z`$. In synchronous co-ordinates the metric is:
$$ds^2=dt^2+A_\mu ^2(t)(dx^\mu )^2$$
(2)
where the metric functions $`A_1(t),A_2(t),A_3(t)`$ are functions of the time co-ordinate only (Greek indices take the space values $`1,2,3`$ and Latin indices the space-time values $`0,1,2,3`$). When two of the functions $`A_\mu (t)`$ are equal (e.g. $`A_2=A_3`$) the Bianchi I space-times reduce to the important class of plane symmetric space-times (a special class of the Locally Rotational Symmetric space-times ) which admit a $`G_4`$ group of isometries acting multiply transitively on the spacelike hypersurfaces of homogeneity generated by the vectors $`\xi _1,\xi _2,\xi _3`$ and $`\xi _4=x^2_3x^3_2`$. In this paper we are interested only in proper diagonal Bianchi I space-times (which in the following will be referred for convenience simply as Bianchi I space-times), hence all metric functions are assumed to be different and the dimension of the group of isometries acting on the spacelike hypersurfaces is three.
A general Bianchi I space-time does not admit a given collineation. The demand that it does, acts like a โselection ruleโ by selecting those Bianchi I space-times whose metric functions $`A_\mu (t)`$ satisfy a certain set of differential equations or algebraic conditions depending on the collineation. These conditions do not necessarily have a solution. For example Coley and Tupper have determined all space-times admitting an ACV. It is easy to check that no (proper) Bianchi I space-time belongs to these space-times. In fact it can be shown that the demand that a Bianchi I space-time admits an ACV leads to the conditions $`A_3(t)=`$const. and $`A_1(t)=A_2(t)`$ i.e. the plane symmetric case.
Although Bianchi I space-times are important in the study of anisotropies and they have served as the basis for this study, it appears that their collineations have not been considered in the literature. For example even at the level of proper conformal symmetries, the CKV found by Maartens and Mellin is a CKV in an LRS space-time and not in a Bianchi I space-time. Perhaps this is due to the fact that the direct solution of the collineation equations is difficult. However there are many general results due to Hall and his co-workers - which will be referred subsequently as they are required - which make possible the determination of the collineations and the corresponding Bianchi I space-times that admit them without solving any difficult differential equations.
In Section II we determine all Bianchi I space-times which admit (proper or not) CKVs. We show that the only Bianchi I space-times which admit a proper HVF are the Kasner-type space-times. In Section III we study the smooth RCs and show that, provided that the Ricci tensor is non-degenerate, there are four families of (proper) Bianchi I space-times admitting smooth RCs. The metrics of these families are determined up to a set of algebraic conditions among the metric functions whereas the corresponding collineation vectors are computed in terms of a set of constant parameters. In Section IV we study the smooth CCs of Bianchi I space-times and show that, assuming the non degeneracy of the Ricci tensor, there are no Bianchi I space-times which admit proper CCs. In Section V we consider the KSS symmetry and determine all Bianchi I space-times which admit a KSS. A particular result which generalises the previous result concerning the HVF, is that the only Bianchi I space-times admitting a proper KSS (of the second kind) are the Kasner-type space-times. In Section VI we discuss the physical implications of these results and examine the compatibility of the physical assumptions on the type of the matter (perfect fluid, electromagnetic field etc.) with the various types of symmetry. An apparently new Bianchi I viscous fluid solution is found, with non zero bulk viscous stress. The model begins with a big-bang and isotropises at late times tending to a Robertson-Walker model (de Sitter Universe).
## 2 Conformal symmetries
To determine the Bianchi I space-times which admit CKVs we use the theorem of Defrise-Carter which has been reconsidered and improved by Hall and Steele . This Theorem concerns the reduction of the conformal algebra $`๐ข`$ of a metric, to the Killing/Homothetic algebra of a (globally defined) conformally related metric. Hall has shown that it is not always possible to find a conformal scaling required by Defrise-Carter, but Hall and Steele showed that the local result can be regained if one imposes at each space-time point the restrictions (a) space-time has the same Petrov type and (b) the dimension of $`๐ข`$ is constant. However, if the Petrov type is I,D,II at a point these restrictions are not necessary. By a direct computation of the Weyl tensor of the general Bianchi I metric (1) we find that the Petrov type is either I (or its degeneracy type D which corresponds to the LRS case which we ignore in this paper) or type O (conformally flat) (in fact all Bianchi type space-times of class A are Petrov type I or its specialisations ). This means that we have two cases to consider i.e. conformally flat and non conformally flat space-times.
### 2.1 Bianchi I space-times of Petrov type I
It is known that the maximum dimension of the conformal algebra of a space-times of Petrov type is I or II is four . Hence the Bianchi I space-time admits at most exactly one proper CKV $`๐`$ (the properness is assured provided that $`Y^00`$) and consequently the four dimensional conformal algebra $`๐ข_4=\{\xi _1,\xi _2,\xi _3,๐\}`$. From the Defrise-Carter theorem it follows that there exists a smooth function $`U(x^a)`$ such that $`๐ข_4`$ restricts to a Lie algebra of KVs for the metric $`d\widehat{s}^2=U^2(x^a)ds^2`$. Because the vectors $`\{\xi _1,\xi _2,\xi _3\}`$ are KVs for both metrics we deduce that $`U(x^a)=U(t)`$ and the type of space-time (i.e. Bianchi I) is retained. Hence the problem of determining the CKVs of Bianchi I space-times is reduced to the determination of the extra KV.
Assuming $`๐=Y^\tau (x^i)_\tau +Y^\mu (x^i)_\mu `$ where $`\tau =U(t)๐t`$ Jacobi identities and Killing equations imply the relations:
$$๐=_\tau +ax_x+by_y+cz_z$$
(3)
$$\widehat{A}_1(\tau )=U(\tau )A_1(\tau )=e^{a\tau },\widehat{A}_2(\tau )=U(\tau )A_2(\tau )=e^{b\tau }\widehat{A}_3(\tau )=U(\tau )A_3(\tau )=e^{c\tau }$$
(4)
where $`abc`$ are integration constants such that at least two are non-zero (otherwise space-time reduces to an LRS space-time). The commutators of the extra KV $`๐`$ with the standard KVs $`\xi _\mu `$ are:
$$[\xi _1,๐]=a\xi _1,[\xi _2,๐]=b\xi _2,[\xi _3,๐]=c\xi _3.$$
In conclusion we have the following result:
All Bianchi I metrics which admit a CKV are $`(abc)`$:
$$ds^2=dt^2+A_1^2(t)\left[dx^2+e^{2(ab)L(t)}dy^2+e^{2(ac)L(t)}dz^2\right]$$
(5)
where:
$$A_1(t)=\frac{1}{U(t)}e^{a{\scriptscriptstyle U(t)๐t}}.$$
(6)
The CKV is given by:
$$๐=\frac{1}{U(t)}_t+ax_x+by_y+cz_z$$
(7)
and has conformal factor:
$$\varphi (๐)=a+\frac{1}{U(t)}\left[\mathrm{ln}\left|A_1(t)\right|\right]_{,t}.$$
(8)
In terms of the time coordinate $`\tau `$ this metric is:
$$ds^2=\frac{1}{U^2(\tau )}\left[d\tau ^2+e^{2a\tau }dx^2+e^{2b\tau }dy^2+e^{2c\tau }dz^2\right]$$
(9)
It is now easy to determine all Bianchi I space-times which admit (one) proper HVF. Indeed setting $`\varphi (๐)=`$const. ($`0`$) and using (8) we find $`U(t)=\frac{1}{\varphi t}`$ from which it follows (ignoring some unimportant integration constants):
###### Proposition 1
The only Bianchi I space-times which admit proper HVF are the Kasner-type space-times given by $`(abc)`$:
$$ds^2=dt^2+t^{2\frac{\varphi a}{\varphi }}dx^2+t^{2\frac{\varphi b}{\varphi }}dy^2+t^{2\frac{\varphi c}{\varphi }}dz^2.$$
(10)
The HVF is:
$$๐=\varphi t_t+ax_x+by_y+cz_z$$
(11)
and has homothetic factor $`\varphi `$.
We remark that the same result can be recovered from the reduction of the proper RCs which will be determined in Section III.
### 2.2 Bianchi I space-times of Petrov type O
The necessary and sufficient condition for conformal flatness is the vanishing of the Weyl tensor $`C_{abcd}`$. By solving directly the equations $`C_{abcd}=0`$ it can be shown that there exist only two families of conformally flat Bianchi I metrics (we ignore the case of the Friedmann-Robertson-Walker space-time) given by:
$$ds_1^2=A_3^2(\tau )ds_{RT}^2$$
(12)
$$ds_2^2=A_3^2(\tau )ds_{ART}^2.$$
(13)
The metrics $`ds_{RT}^2,ds_{ART}^2`$ have been found previously by Rebouรงas-Tiomno and Rebouรงas-Teixeira respectively and are 1+3 (globally) decomposable space-times whose 3-spaces are spaces of constant curvature. In synchronous co-ordinates they are:
$$ds_{RT}^2=dz^2d\tau ^2+\mathrm{cos}^2(\frac{\tau }{a})dy^2+\mathrm{sin}^2(\frac{\tau }{a})dx^2$$
(14)
$$ds_{ART}^2=dz^2d\tau ^2+\mathrm{cosh}^2(\frac{\tau }{a})dy^2+\mathrm{sinh}^2(\frac{\tau }{a})dx^2$$
(15)
where $`d\tau =\frac{dt}{A_3(t)}`$. Each metric admits 15 CKVs which can be determined using standard techniques . In concise notation these vectors are (we ignore the KVs $`\xi _\mu `$ which constitute the $`G_3`$) ($`k=0,1,2,3`$ and $`\alpha =0,1,2`$):
KVs
$$\xi _{k+4}=\left\{s_\pm (y,a)\left[\delta _k^0c_+(x,a)+\delta _k^1s_+(x,a)\right]+c_\pm (y,a)\left[\delta _k^2c_+(x,a)+\delta _k^3s_+(x,a)\right]\right\}_\tau +$$
$$+\frac{s_{}(\tau ,a)}{c_{}(\tau ,a)}\left\{c_\pm (y,a)\left[\delta _k^0c_+(x,a)+\delta _k^1s_+(x,a)\right]\pm s_\pm (y,a)\left[\delta _k^2c_+(x,a)+\delta _k^3s_+(x,a)\right]\right\}_y$$
$$\frac{c_{}(\tau ,a)}{s_{}(\tau ,a)}\left\{s_\pm (y,a)\left[\delta _k^0s_+(x,a)+\delta _k^1c_+(x,a)\right]+c_\pm (y,a)\left[\delta _k^2s_+(x,a)+\delta _k^3c_+(x,a)\right]\right\}_x$$
(16)
CKVs
$$X_{(k)\alpha }=\pm a^2B_{k,\alpha }X_{(k)3}=a^2B_{k,3}$$
(17)
$$\varphi (๐_{(k)})=B_k$$
(18)
$$X_{(k+4)\alpha }=\pm a^2\mathrm{\Gamma }_{k,\alpha }X_{(k+4)3}=a^2\mathrm{\Gamma }_{k,3}$$
(19)
$$\varphi (๐_{(k+4)})=\mathrm{\Gamma }_k$$
(20)
where:
$$B_k=c_{}(\tau ,a)\{c_\pm (y,a)[s_{}(z,a),c_{}(z,a)],s_\pm (y,a)[s_{}(z,a),c_{}(z,a)]\}$$
(21)
$$\mathrm{\Gamma }_k=s_{}(\tau ,a)\{c_+(x,a)[s_{}(z,a),c_{}(z,a)],s_+(x,a)[s_{}(z,a),c_{}(z,a)]\}.$$
(22)
and the following conventions have been used:
1. The upper sign corresponds to RT space-time and the lower sign to ART space-time.
2. The functions $`s_{}(w,a),c_{}(w,a)`$ are defined as follows:
$$(c_+(w,a),c_{}(w,a))=(\mathrm{cosh}(\frac{w}{a}),\mathrm{cos}(\frac{w}{a}))$$
(23)
$$(s_+(w,a),s_{}(w,a))=(\mathrm{sinh}(\frac{w}{a}),\mathrm{sin}(\frac{w}{a}))$$
(24)
The Bianchi I metrics (12), (13) have the same CKVs with conformal factors $`\psi (\xi _{k+4})=\xi _{k+4}(\mathrm{ln}A_3)`$ and $`\psi (๐_A)=๐_A(\mathrm{ln}A_3)+\varphi (๐_A)`$ ($`A=1,2,\mathrm{},8`$). It follows that conformally flat Bianchi I space-times do not admit HVFs. Furthermore if we enforce them to admit an extra KV they reduce to the RT and the ART space-times which admit seven KVs.
## 3 Ricci Collineations
A RC $`๐=X^a_a`$ is defined by the condition:
$$_๐R_{ab}=R_{ab,c}X^c+R_{ac}X_{,b}^c+R_{bc}X_{,a}^c=0.$$
(25)
For RCs there do not exist theorems of equal power to the Theorem of Defrise-Carter and one has to solve directly the differential equations (25). However there do exist some general results available which are due to Hall and are summarised in the following statement :
If the Ricci tensor is of rank 4, at every point of the space-time manifold, then the smooth ($`C^2`$ is enough) RCs form a Lie algebra of smooth vector fields whose dimension is $`10`$ and $`9`$. This Lie algebra contains the proper RCs and their degeneracies.
It has been pointed out by Hall and his co-workers that the assumption on the order of the Ricci tensor is important. Indeed let us assume that the order of the Ricci tensor of a Bianchi I space-time is 3. Then due to the fact that the Ricci tensor in Bianchi I space-times is diagonal one of its components must vanish, the $`R_{11}=0`$ say. This is equivalent to $`R_{ab}\xi _1^b=0`$ where $`\xi _1=_x`$. Consider the vector field $`๐_1=f(x^a)_x`$ where $`f(x^a)`$ is an arbitrary (but smooth) function of its arguments. It is easy to show that $`_{๐_1}R_{ab}=0`$ so that $`๐_1`$ is a Ricci collineation. Due to the arbitrariness of the function $`f(x^a)`$ these Bianchi I space-times admit infinitely many smooth RCs a result that does not help us in any useful or significant way in their study.
In the following we consider smooth RCs and we assume that $`R_{ab}`$ is non-degenerate ($`detR_{ab}0`$). Equation (25) gives the following set of 10 differential equations (no summation over the indices $`\mu ,\nu ,\rho =1,2,3;\mu \nu \rho `$):
$$[00]\text{ }R_{00,0}X^0+2R_{00}X_{,0}^0=0$$
(26)
$$[0\mu ]\text{ }R_{00}X_{,\mu }^0+R_{\mu \mu }X_{,0}^\mu =0$$
(27)
$$[\mu \mu ]\text{ }R_{\mu \mu ,0}X^0+2R_{\mu \mu }X_{,\mu }^\mu =0$$
(28)
$$[\mu \nu ]\text{ }R_{\mu \mu }X_{,\nu }^\mu +R_{\nu \nu }X_{,\mu }^\nu =0$$
(29)
where a comma denotes partial differentiation w.r.t. following index co-ordinate. For convenience we set $`R_{00}R_0,R_{\mu \mu }R_\mu `$. Equation (26) is solved immediately to give:
$$X^0=\frac{m(x^\beta )}{\sqrt{|R_0|}}$$
(30)
where $`m(x^\beta )`$ is a smooth function of the spatial co-ordinates. Using (30) we rewrite the remaining equations (27)-(29) in the form:
$$X_{,0}^\mu =ฯต_0\frac{\sqrt{\left|R_0\right|}}{R_\mu }m_{,\mu }$$
(31)
$$X_{,\mu }^\mu =\frac{\left(\mathrm{ln}\left|R_\mu \right|\right)_{,0}}{2\sqrt{\left|R_0\right|}}m$$
(32)
$$R_\mu X_{,\nu }^\mu +R_\nu X_{,\mu }^\nu =0$$
(33)
where $`\mu \nu `$ and $`ฯต_0`$ is the sign of the component $`R_0`$.
Differentiating equation (33) w.r.t. $`x^\rho `$ we obtain:
$$R_\mu X_{,\nu \rho }^\mu +R_\nu X_{,\mu \rho }^\nu =0.$$
(34)
Rewriting (33) for the indices $`\mu ,\rho `$, differentiating w.r.t. $`x^\nu `$ and subtracting from (34) we obtain:
$$R_\nu X_{,\mu \rho }^\nu R_\rho X_{,\nu \mu }^\rho =0.$$
(35)
Writing (33) for the indices $`\nu ,\rho `$, differentiating w.r.t. $`x^\mu `$ and adding to (35) ($`R_\mu 0`$) we get:
$$X_{,\mu \rho }^\nu =0\text{(}\mu \nu \rho \text{).}$$
(36)
Differentiating (31) w.r.t. $`x^\mu `$ and (32) w.r.t. $`x^0`$ we find:
$$\frac{R_\mu }{\sqrt{\left|R_0\right|}}\left[\frac{\left(\mathrm{ln}\left|R_\mu \right|\right)_{,0}}{2\sqrt{\left|R_0\right|}}\right]_{,0}=a_\mu $$
(37)
$$m_{,\mu \mu }=ฯต_0a_\mu m$$
(38)
where $`a_\mu `$ are arbitrary constants.
Equations (31)-(33) and (36)-(38) constitute a set of differential equations in the variables $`(R_\mu ,X^\mu ,m)`$, which can be solved in terms of $`R_0`$ and some integration constants. These constants are constrained by a set of algebraic equations involving the spatial components of the Ricci tensor and essentially determine the dimension of the algebra of the RCs.
The complete set of the solutions consists of five main cases which are summarized in TABLE I together with the corresponding algebraic constraints and the dimension of the resulting algebra.
To find the RCs we note that in all cases the component $`X^0`$ is given by (30) and it is determined in terms of the function $`m(x^\alpha ).`$ Concerning the spatial components these are obtained from the following formulas taking into consideration the fourth column of TABLE I.
Case A
$$X_I^\mu =c_\mu \left\{dx^\mu x^\mu \underset{\nu \mu }{}D_\nu x^\nu \frac{D_\mu }{2}\left[(x^\mu )^2\underset{\nu \mu }{}\left(\frac{C_\mu }{C_\nu }\right)(x^\nu )^2\frac{ฯต_0}{c_\mu ^2R_\mu }\right]\right\}+\underset{\nu \mu }{}B_\nu ^\mu x^\nu $$
(39)
Case B ($`A=2,3)`$
$$X_{II}^A=c_A\left\{dx^Ax^AD_Bx^B\frac{D_A}{2}\left[(x^A)^2\left(\frac{C_A}{C_B}\right)(x^B)^2\frac{ฯต_0}{c_A^2R_A}\right]\right\}+\mathrm{\Lambda }_B^Ax^B.$$
(40)
and $`X_{II}^1=0`$.
Case Ca
$$X_{III}^1=\frac{ฯต_0m_{,1}}{R_1}\sqrt{\left|R_0\right|}dt+b_2^1x^2,X_{III}^2=\frac{ฯต_0m_{,2}}{R_2}\sqrt{\left|R_0\right|}dt+b_1^2x^1$$
(41)
$$X_{III}^3=\frac{ฯต_0m_{,3}}{2a\sqrt{\left|R_0\right|}}\left(\mathrm{ln}\left|R_3\right|\right)_{,t}$$
(42)
Case Cb
$$X_{IV}^1=b_2^1x^2,X_{IV}^2=b_1^2x^1$$
(43)
$$X_{IV}^3=\frac{ฯต_0D_3}{2c_3R_3}c_3\left[D_3\frac{(x^3)^2}{2}+dx^3\right]$$
(44)
Case D
$$X_V^\mu =\underset{\nu \mu }{}b_\nu ^\mu x^\nu +f^\mu (x^0)$$
(45)
where:
$$f^\mu (x^0)=\frac{D_\mu }{R_\mu }\sqrt{\left|R_0\right|}๐x^0.$$
(46)
We collect the above results in the following:
###### Proposition 2
The proper smooth RCs in Bianchi I space-times can be considered in four sets depending on the constancy of the spatial Ricci tensor components. The first set (case A with $`R_{\mu ,0}0`$ for all $`\mu =1,2,3`$) contains three families of smooth RCs consisting of either one, two or seven RCs defined by the vector $`๐_I`$ given by (39). The second set (case B with one $`R_{\mu ,0}=0`$) consists of two families of one and four RCs defined by the vector $`๐_{II}`$ given by (40). The third set (case C with two $`R_{\mu ,0}=0)`$ consists of five families with three, three, three, seven and three RCs are given by the vector fields $`๐_{III},๐_{IV}`$ defined in (41)-(42), (43)-(44). Finally the last set (case D all $`R_{\mu ,0}=0)`$ contains one family of seven RCs given by the vector field $`๐_V`$ defined in (46). In each family of RCs the spatial Ricci tensor components are given in terms of the time component $`R_0`$ and the coefficients of the vector fields are constrained with the spatial components of the Ricci tensor via algebraic conditions.
We note that RCs do not fix the metric up to a set of constants as the CKVs (and the lower symmetries) do but instead they impose algebraic conditions on the metric functions. Thus in general one should expect many families of Bianchi I space-times admitting RCs.
## 4 Curvature Collineations
Curvature collineations are necessarily Ricci Collineations and one is possible to determine them from the results of the last section. The easiest way to do this would appear to use algebraic computing algorithms and compute directly the Lie derivative of the curvature tensor for the RCs found in the last section. However this is not so obvious because although we know the RC we do not know the metric functions. Hence, in general, one expects to arrive at a system of differential equations among the metric functions $`A_\mu (t)`$ whose solution will give the answer.
However a study of the relevant literature shows that there are enough general results which allow one to determine the CCs without solving any differential equations. The CCs have many of the pathologies of RCs. For example for any positive integer $`k`$ there are CCs which are $`C^k`$ but not $`C^{k+1}.`$ Furthermore they may form an infinite dimensional vector space which is not a Lie algebra under the usual Lie bracket operation. However if one considers the $`C^{\mathrm{}}`$ CCs only (loosing in that case the ones that are not smooth) then they do form a Lie algebra which is a subalgebra of the (smooth) RCs algebra .
For the determination of CCs in Bianchi I space-times it is enough to use the following result from an early work of Hall :
If the curvature components are such that at every point of a space-time M the only solution of the equation $`R_{abcd}k^d=0`$ is $`k^d=0`$ then every CC on M is a HVF.
Let us assume that in a Bianchi I space-time the equation $`R_{abcd}k^d=0`$ has a solution $`k^a0`$. Then it follows that equation $`R_{ab}k^d=0`$ admits a non-vanishing solution which is impossible because $`R_{ab}`$ is non-degenerate. Hence according to the above statement all $`C^{\mathrm{}}`$ CC in Bianchi I space-times are HVFs or, equivalently there are no Bianchi I space-times (with non-degenerate Ricci tensor) which admit proper CCs.
## 5 Kinematic Self Similarities
As it has been mentioned in the Introduction, Kinematic Self Similarities are not geometric symmetries (i.e. collineations). They are kinematic symmetries/constraints which involve the 4-velocity of the fluid (or in empty space-times a timelike unit vector field) defined by the conditions:
$$_Xu_a=\alpha u_a_Xh_{ab}=2\delta h_{ab}$$
(47)
where $`h_{ab}=g_{ab}+u_au_b`$ projects normally to $`u_a`$ and $`\alpha ,\delta `$ are constants.
Kinematic self similarities have been studied by Sintes who determined all LRS perfect fluid space-times which admit a KSS. Our aim in this Section is to determine the KSS of (proper) Bianchi I metrics without any restriction on the type of the fluid except that we assume that the fluid 4-velocity is orthogonal to the group orbits i.e. $`u^a=\delta _0^a`$. This assumption enforces the commutator of a KSS $`๐`$ with the three KVs $`\xi _\mu `$ to be a KV. Hence we write:
$$[\xi _\mu ,๐]=X_{,\mu }^a_a=a_\mu ^\nu \xi _\nu $$
(48)
where $`a_\mu ^\nu `$ are constants. Integrating we find:
$$X^0=X^0(t)\text{and}X^\mu =a_\nu ^\mu x^\nu +f^\mu (t)$$
(49)
where $`f^\mu (t)`$ are arbitrary smooth functions of their argument. The first of equations (47) gives:
$$X^0=\alpha t+\beta $$
where $`\beta `$ is an integration constant and (without loss of generality) $`f^\mu (t)=`$const.$`=0`$. The second equation of (47) gives the following conditions among the metric functions:
$$a_\nu ^\mu (A_\mu )^2+a_\mu ^\nu (A_\nu )^2=0$$
(50)
$$a_\mu ^\mu +(\alpha t+\beta )\frac{d(\mathrm{ln}A_\mu )}{dt}=\delta $$
(51)
where $`\mu \nu `$ and a dot denotes differentiation w.r.t. $`t`$. Equation (50) means that, in order to avoid the plane symmetric case, we must take $`a_\nu ^\mu =0`$ for $`\mu \nu `$. Therefore we have the following conclusion ($`a_\mu ^\mu a_\mu `$):
###### Proposition 3
The Bianchi I space-times whose metric functions satisfy the relation ($`a_\mu ฯตR`$):
$$a_\mu +(\alpha t+\beta )\frac{d(\mathrm{ln}A_\mu )}{dt}=\delta $$
(52)
admit the proper KSS:
$$๐=(\alpha t+\beta )_t+a_1x_x+a_2y_y+a_3z_x.$$
(53)
In view of equation (52) we have two distinct cases to consider, namely $`\alpha =0`$ (type zero) and $`\alpha 0`$.
Case $`\alpha =0`$ $`(\delta 0)`$.
For $`\beta =0`$ equation (52) is trivially satisfied i.e. all Bianchi I space-times admit the (zeroth kind) KSS:
$$๐=\delta (x_x+y_y+z_x).$$
(54)
For $`\beta 0`$ we have:
$$A_\mu (t)=e^{\frac{\delta a_\mu }{\beta }t}.$$
(55)
It is easy to show that in this case space-time admits the KV:
$$๐=_t+\frac{a_1\delta }{\beta }x_x+\frac{a_2\delta }{\beta }y_y+\frac{a_3\delta }{\beta }z_x$$
(56)
and becomes homogeneous. (The KV (56) can also be found from the reduction of the results of Section II).
Case $`\alpha 0`$
In this case the solution of (52) is:
$$A_\mu (t)=(\alpha t+\beta )^{\frac{\delta a_\mu }{\alpha }}$$
(57)
and leads to the conclusion that:
The only Bianchi I space-times with co-moving fluid which admit a KSS of the second kind are the Kasner type space-times.
We note that the Bianchi I space-times with metric functions given by equation (57) also admit the HVF:
$$\alpha _t+(\alpha +a_1\delta )x_x+(\alpha +a_2\delta )y_y+(\alpha +a_3\delta )z_x$$
(58)
with homothetic factor $`\alpha `$ (see Section II).
## 6 Discussion
Working with purely geometric methods we have succeeded to determine all (proper and diagonal) Bianchi I space-times which admit certain (and the most important) collineations. In many cases the explicit form of the metrics has been given (CKVs, HVFs, KSS), in others the metrics are defined up to a set of conditions on the metric functions (RCs) and finally it has been shown that there are not Bianchi I space-times which admit CCs and ACVs. In all cases the collineation vector has been determined (whenever it exists).
In order to establish the physical significance of these general geometrical results we address the following questions for each type of collineation:
1. Are there any Bianchi I metrics among the ones selected by one of the collineations considered, which satisfy the energy conditions, so that they can be used as potential space-time metrics?
2. If there are, are the known Bianchi I solutions among these solutions?
In the following we take the cosmological constant $`\mathrm{\Lambda }=0`$.
### 6.1 The case of CKVs
Of interest is only the non-conformally flat metrics (6) (or (7)) which admit the CKV (9) with conformal factor (8).
One general result is that the CKV is inheriting, that is, the fluid flow lines are preserved under Lie transport along the CKV.
Concerning the dynamical results we consider three main cases: perfect fluids, non null Einstein-Maxwell solutions and imperfect fluid solutions (the case of null Einstein-Maxwell field is excluded because the Segrรฉ type of Bianchi I space-times is $`[1,111]`$ or degeneracies of this type).
Perfect Fluid Solutions
In this case the Segrรฉ type of space-time is $`[1,(111)]`$ and all spatial eigenvalues $`\lambda _\mu `$ are equal. Moreover $`\lambda _\mu =G_\mu ^\mu `$ where $`G_{ab}`$ is the Einstein tensor. Considering $`\lambda _2=\lambda _3`$ and using (8) we find $`U=\frac{1}{Bt}`$ where $`B=\frac{a+b+c}{2}`$. Finally replacing $`U(t)`$ in the line element (6) we find that the resulting space-time is a Kasner type space-time and the CKV reduces to a HVF in agreement with the general result that orthogonal spatially homogeneous perfect fluid space-times do not admit any inheriting proper CKV. Most of the known Bianchi I solutions concern perfect fluid solutions. None of these solutions admit a CKV.
Einstein Maxwell solutions
For these fields the Segrรฉ type of the Einstein tensor is $`[(1,1)(11)]`$ hence $`\lambda _2=\lambda _3`$ and $`\lambda _0=\lambda _1`$. The first equality implies again that the metric reduces to a Kasner type metric and in fact to a vacuum solution (because if we force a Kasner type metric to represent a (necessarily non null) Einstein-Maxwell field it reduces to a vacuum solution).
###### Proposition 4
There do not exist Bianchi I (non-null) Einstein-Maxwell space-times which admit a CKV or a HVF.
The two Bianchi I solutions with electromagnetic field found by Datta and by Rosen do not admit a CKV or a HVF.
Anisotropic fluid solutions
The above results indicate that the Bianchi I metrics (6) can represent only anisotropic fluid space-times. Recently anisotropic fluid Bianchi I cosmological models have been investigated extensively using a dynamical system approach and the truncated Israel-Stewart theory of irreversible thermodynamics. It has been found that in these models, anisotropic stress leads to models which violate the weak energy condition, thus they are unphysical or they lead to the creation of a periodic orbit .
In order to find one such solution which will be physically viable we consider the following two restrictions:
1. $`c=0`$ and
2. The algebraic type of matter (equivalently Einstein) tensor is $`[(1,1)11].`$
Setting $`\lambda _0=\lambda _3`$ we obtain the condition:
$$2(\mathrm{ln}M)_{,\tau \tau }2\left[(\mathrm{ln}M)_{,\tau }\right]^2+b^2+a^2=0$$
(59)
whose solution is:
$$M(\tau )=\frac{1}{\mathrm{sinh}k\tau }\text{ },\text{ }\frac{1}{\mathrm{cosh}k\tau }$$
(60)
where $`k^2=\frac{a^2+b^2}{2}`$. We keep the solution $`M(\tau )=\frac{1}{\mathrm{sinh}k\tau }`$ because the other violates all energy conditions. This gives $`U(\tau )=\mathrm{sinh}k\tau `$ or $`U(t)=\mathrm{sinh}^1kt`$ and finally we obtain the metric:
$$ds^2=dt^2+\mathrm{sinh}^{2\frac{ka}{k}}\frac{kt}{2}\mathrm{cosh}^{2\frac{k+a}{k}}\frac{kt}{2}dx^2+\mathrm{sinh}^{2\frac{kb}{k}}\frac{kt}{2}\mathrm{cosh}^{2\frac{k+b}{k}}\frac{kt}{2}dy^2+\mathrm{sinh}^2ktdz^2.$$
(61)
This new Bianchi I space-time describes a viscous fluid and satisfies the weak and the dominant energy conditions (a description of these energy conditions in terms of the eigenvalues of the stress-energy tensor is given in the Appendix) provided $`ab>0`$ and $`ka<0`$. The strong energy condition is violated. It also admits the proper CKV $`๐=\mathrm{sinh}kt_t+ax_x+by_y`$ with conformal factor $`\varphi (๐)=k\mathrm{cosh}kt`$.
To study the physics of the new solution we consider the stress-energy tensor $`T_{ab}`$ and using the standard Eckart theory we write:
$$T_{ab}=\mu u_au_b+(\overline{p}\zeta \theta )h_{ab}2\eta \sigma _{ab}$$
(62)
where $`\zeta ,\eta 0`$ are the bulk and the shear viscosity coefficients and $`\overline{p}`$ is the isotropic pressure in the absence of dissipate processes i.e. $`\zeta =0`$ (equilibrium state). It is easy to show that vanishing of $`\zeta `$ together with a linear barotropic equation of state $`\overline{p}=(\gamma 1)\mu `$ (where $`\gamma [1,2]`$) lead to the condition $`a=b,`$ which violates the weak energy condition ($`ab<0`$). Thus we restrict our study to the case where $`\zeta 0`$ which is of cosmological interest. For example inflation driven by a viscous fluid necessarily involves bulk viscous stress. Moreover cosmological models which include viscosity can be used in an attempt to interpret the observed highly isotropic matter distribution. In fact it has been shown that viscosity plays a significant role in the isotropisation of the cosmological models.
Using standard methods we find for the kinematic and the dynamic variables of the model:
$`\mu `$ $`=`$ $`{\displaystyle \frac{3k^2\mathrm{cosh}^2kt2k(b+a)\mathrm{cosh}kt+ab}{\mathrm{sinh}^2kt}}`$
$`\zeta `$ $`=`$ $`{\displaystyle \frac{1}{\theta }}\left[\mu +\overline{p}+{\displaystyle \frac{2k(a+b)\mathrm{cosh}kt2(k^2+ab)}{3\mathrm{sinh}^2kt}}\right]`$ (63)
$`\theta `$ $`=`$ $`{\displaystyle \frac{3k\mathrm{cosh}kt(b+a)}{\mathrm{sinh}kt}}`$
$`\sigma _{11}`$ $`=`$ $`{\displaystyle \frac{(b2a)}{6}}\mathrm{sinh}^{\frac{k2a}{k}}{\displaystyle \frac{kt}{2}}\mathrm{cosh}^{\frac{k+2a}{k}}{\displaystyle \frac{kt}{2}}`$
$`\sigma _{22}`$ $`=`$ $`{\displaystyle \frac{(a2b)}{6}}\mathrm{sinh}^{\frac{k2b}{k}}{\displaystyle \frac{kt}{2}}\mathrm{cosh}^{\frac{k+2b}{k}}{\displaystyle \frac{kt}{2}}`$ (64)
$`\sigma _{33}`$ $`=`$ $`{\displaystyle \frac{(a+b)}{6}}\mathrm{sinh}kt`$
$$\sigma ^2=\frac{a^2ab+b^2}{3\mathrm{sinh}^2kt}$$
(65)
where $`2\sigma ^2=\sigma _{ab}\sigma ^{ab}`$ and:
$$\eta =\frac{a+b}{2\mathrm{sinh}kt}k\mathrm{coth}kt.$$
(66)
The explicit computation of the bulk viscosity $`\zeta `$ requires the adoption of a specific equation of state in order to guarantee that $`\zeta `$ is positive definite. However the simple choice $`a+b>0`$ and $`k<0`$ ensures that $`\zeta 0`$ (since $`\theta <0`$) and $`\eta >0`$ provided that $`\overline{p}p_{eff}\overline{p}\zeta \theta `$.
Concerning the asymptotic behavior of the model we have $`\underset{t\mathrm{}}{lim}\sigma =0`$ provided that $`k<0`$ and the model isotropizes at late times. Furthermore $`\underset{t\mathrm{}}{lim}\theta =3k`$ , $`\underset{t\mathrm{}}{lim}R=12k^2,`$ $`(\mu +p_{eff})_t\mathrm{}=0`$ hence the model corresponds to the flat FRW space-time i.e. the de Sitter universe. This also follows directly from the metric (61) if we consider the limit $`t\mathrm{}`$. In this limit the CKV $`๐`$ degenerates to a KV. In addition using (63) it can be shown that there is a cosmological singularity of Kasner type at a finite time in the past i.e. there is a $`t=t_0`$ at which the energy density vanishes, whereas the initial singularity occurs at $`t=0`$.
### 6.2 The case of Ricci Collineations
There are four families of Bianchi I space-times which admit RCs and furthermore the metric in these families is fixed only up to a set of algebraic conditions. Due to this generality we are obliged to consider again special cases and the best choice is perfect fluid solutions.
The algebraic type of the Ricci tensor for a perfect fluid is $`[1,(111)]`$ which implies the condition:
$$\frac{R_{11}}{A_1^2}=\frac{R_{22}}{A_2^2}=\frac{R_{33}}{A_3^2}.$$
(67)
This immediately excludes the last three families (B,Ca,Cb,D) of TABLE I and we are left with family A only. From the first column of TABLE I we read $`R_\mu =C_\mu e^{{\scriptscriptstyle 2c_\mu \sqrt{\left|R_0\right|}๐t}}`$ hence (67) imply:
$$\frac{C_1}{C_2}e^{2(c_1c_2){\scriptscriptstyle \sqrt{\left|R_0\right|}๐t}}=\frac{A_1^2}{A_2^2}\text{and}\frac{C_1}{C_3}e^{2(c_1c_3){\scriptscriptstyle \sqrt{\left|R_0\right|}๐t}}=\frac{A_1^2}{A_3^2}.$$
(68)
In order to avoid the FRW metric ($`A_1A_2A_3)`$ and the plane symmetric metric (e.g. $`A_1A_2`$) we demand $`c_1c_2c_3`$. Then from the third column of TABLE I we have that only case A<sub>2</sub> survives and furthermore $`D_\mu =b_\nu ^\mu =0`$. Setting the constant $`d=1`$ we find from (39):
$$๐=\frac{1}{\sqrt{R_{00}}}_0c_1x^1_1c_2x^2_2c_3x^3_3.$$
(69)
We conclude that all perfect fluid Bianchi I space-times which satisfy (68) admit a RC of the form (69).
As far as we aware all existing perfect fluid solutions in Bianchi I space-times concern perfect fluids with a linear barotropic equation of state $`p=(\gamma 1)\mu `$ ($`\gamma [1,2]`$). In order to compare these solutions with the above perfect fluid solutions we assume that the later also satisfy a linear barotropic equation of state. Then from the field equations we obtain:
$$R_{ab}=\mu \left[\frac{(2\gamma )}{2}g_{ab}+\gamma u_au_b\right].$$
(70)
(The value $`\gamma =2`$ is excluded because then the Ricci tensor becomes degenerated). The 00 conservation equation gives:
$$\mu _{,t}+\gamma \mu \theta =0$$
(71)
where $`\theta =u_{;a}^a`$ is the expansion of the fluid. Introducing the scale factor or โmean radiusโ $`S^3=A_1A_2A_3`$ we find $`\theta =\left(\mathrm{ln}S^3\right)_{,t}`$ and the solution of equation (71) is:
$$\mu =\frac{M}{S^{3\gamma }}$$
(72)
where $`M=`$ constant. Using (68), (70) and (72) we find:
$$\mu =\frac{\text{const.}}{t^2}.$$
(73)
Combining this with (67) and (68) we find $`A_1t^p,A_2t^q,A_3t^r`$ i.e. the resulting Bianchi space-time is a Kasner type space-time which is a contradiction because a Kasner type space-time with a perfect fluid leads to stiff matter i.e. $`\gamma =2`$. Thus we have proved:
###### Proposition 5
All perfect fluid Bianchi I space-times whose metric functions satisfy condition (68) admit a RC of the form (69). In addition perfect fluid Bianchi I space-times with linear barotropic equation of state do not admit proper RCs.
Finally it should be pointed that although spatially homogeneous perfect fluid space-times must satisfy a barotropic equation of state $`p=p(\mu )`$ this equation need not necessarily be linear. However it was proved recently that the asymptotic behaviour of such models is similar to the case of a linear barotropic equation of state .
Acknowledgments
We thank the referee for useful and accurate remarks and suggestions. One of the authors (P.S.A.) was partially supported by the Hellenic Fellowship Foundation (I.K.Y.).
Appendix
It is well known that Bianchi I space-times have zero heat flux and the 4-velocity $`u^a`$ is an eigenvector of the energy momentum tensor. Consequently the only possible algebraic Segrรฉ type of their energy momentum tensor is $`[1,111]`$ and its degeneracies . Furthermore, for this type of energy momentum tensors it has been shown that the energy conditions take the following form :
Weak energy condition
$$\lambda _00\text{ and }\lambda _0+\lambda _\alpha 0$$
(A1)
Dominant energy condition
$$\lambda _00\text{ and }\left|\lambda _\alpha \right|\lambda _0$$
(A2)
Strong energy condition
$$\lambda _00\text{ and }\lambda _0+\underset{\alpha }{}\lambda _\alpha 0$$
(A3)
where $`\lambda _0`$ is the eigenvalue of the timelike eigenvector and $`\lambda _\alpha `$ are the eigenvalues of the spacelike eigenvectors.
TABLE I: The table contains the complete set of solutions of the Ricci Collineation equations. The solutions are specified in terms of $`R_0`$ and some integration constants together with the algebraic constraints (if any) which $`R_\mu `$ must satisfy. The second column contains the spatial components of the Ricci tensor, the third the function $`m(x^\alpha )`$, the fourth column the constraints on the integration constants and the last column the number of proper RCs. The indices $`\mu ,\nu ,\rho =1,2,3`$ , $`\mu \nu \rho `$ , $`A,B=2,3`$ $`AB`$ and there is no summation over repeated indices. The functions $`c_\epsilon (x^3,a),s_\epsilon (x^3,a)`$ are given in (23) and (24).
|
warning/0006/hep-ph0006252.html
|
ar5iv
|
text
|
# HIGHERโTWIST CONTRIBUTIONS TO ๐น_๐ฟ AND ๐นโ FROM INSTANTONSโ
## References
|
warning/0006/gr-qc0006081.html
|
ar5iv
|
text
|
# J. Math. Phys.35, 4184 (1994) Null cone evolution of axisymmetric vacuum spacetimes
## I Introduction
The physical basis of a new algorithm for the evolution of spacetimes has been proposed. This algorithm is based upon the characteristic initial value problem for general relativity, using light cones as the evolution hypersurfaces, rather than the spacelike foliation used in traditional approaches based upon the Cauchy problem. The intimate use of characteristics has particular advantages for the description of gravitational radiation and black hole formation. . The first attempt to carry out numerical evolution based upon this algorithm was only successful in a region outside some sufficiently large worldtube. Near the vertex of the null cone, instabilities arose which destroyed the accuracy of the code. The underlying cause of this instability was too complicated to analyze in the context of general relativity, especially since the numerical analysis of the characteristic initial value problem had not yet been carried out even for the simplest linear axisymmetric systems.
This warranted an investigation of the basic computational properties of the evolution of the flat space scalar wave equation using a null cone initial value formulation . It was found that near the vertex of the cone the Courant-Friedrichs-Lewy (CFL) condition places a stricter limit on the time step than for the case of Cauchy evolution. This insight made possible the development of an extremely efficient marching algorithm for evolving the data on the initial cone by stepping it out from the vertex of the next cone to null infinity (scri) along each angular ray direction. This marching algorithm is based upon a simple identity satisfied by the values of the scalar field at the corners of a parallelogram formed by four null rays. The result was a stable, calibrated, second order accurate global algorithm on a compactified grid. Furthermore, scri played the role of a perfect absorbing boundary so that no radiation was reflected back into the system. This algorithm offers a powerful new approach to generic wave type systems. The basic idea is applicable to any of the hyperbolic systems occurring in physics.
In this paper, we apply the algorithm to the evolution of asymptotically flat, (twist-free) axisymmetric spacetimes. In a Bondi null coordinate system the metric takes the form
$`ds^2`$ $`=`$ $`({\displaystyle \frac{V}{r}}e^{2\beta }U^2r^2e^{2\gamma })du^2+2e^{2\beta }dudr+2Ur^2e^{2\gamma }dud\theta `$ (2)
$`r^2(e^{2\gamma }d\theta ^2+e^{2\gamma }\mathrm{sin}^2\theta d\varphi ^2).`$
The vacuum field equations then decompose into the three hypersurface equations
$$\beta _{,r}=\frac{1}{2}r(\gamma _{,r})^2$$
(3)
$$[r^4e^{2(\gamma \beta )}U_{,r}]_{,r}=2r^2[r^2(\frac{\beta }{r^2})_{,r\theta }\frac{(\mathrm{sin}^2\theta \gamma )_{,r\theta }}{\mathrm{sin}^2\theta }+2\gamma _{,r}\gamma _{,\theta })]$$
(4)
$`V_{,r}`$ $`=`$ $`{\displaystyle \frac{1}{4}}r^4e^{2(\gamma \beta )}(U_{,r})^2+{\displaystyle \frac{(r^4\mathrm{sin}\theta U)_{,r\theta }}{2r^2\mathrm{sin}\theta }}`$ (6)
$`+e^{2(\beta \gamma )}[1{\displaystyle \frac{(\mathrm{sin}\theta \beta _{,\theta })_{,\theta }}{\mathrm{sin}\theta }}+\gamma _{,\theta \theta }+3\mathrm{cot}\theta \gamma _{,\theta }(\beta _{,\theta })^22\gamma _{,\theta }(\gamma _{,\theta }\beta _{,\theta })]`$
and one evolution equation
$`4r(r\gamma )_{,ur}`$ $`=`$ $`[2r\gamma _{,r}Vr^2(2\gamma _{,\theta }U+\mathrm{sin}\theta ({\displaystyle \frac{U}{\mathrm{sin}\theta }})_{,\theta })]_{,r}2r^2{\displaystyle \frac{(\gamma _{,r}U\mathrm{sin}\theta )_{,\theta }}{\mathrm{sin}\theta }}`$ (8)
$`+{\displaystyle \frac{1}{2}}r^4e^{2(\gamma \beta )}(U_{,r})^2+2e^{2(\beta \gamma )}[(\beta _{,\theta })^2+\mathrm{sin}\theta ({\displaystyle \frac{\beta _{,\theta }}{\mathrm{sin}\theta }})_{,\theta }].`$
The initial data consists of $`\gamma `$, which is unconstrained except by smoothness conditions. Because $`\gamma `$ represents a spin-2 field, it must be $`O(\mathrm{sin}^2\theta )`$ near the axis and consist of $`l2`$ spin-2 multipoles.
Here we are interested in the global application of this system when the null hypersurfaces are null cones, although the approach also goes through without major change if the null hypersurfaces emanate from a finite world tube. We require that the null cones have nonsingular vertices which trace out a geodesic worldline $`r=0`$. For the quadrupole terms, this implies the boundary conditions $`\gamma =O(r^2)`$, $`\beta =O(r^4)`$, $`U=O(r)`$ and $`V=r+O(r^3)`$. For higher multipoles, the smoothness conditions can be worked out by referring back to local Minkowski coordinates . As a consequence, $`O(r^n)`$ terms in $`\gamma `$ can contain multipoles with $`2ln`$. Any satisfactory computational algorithm must meet the challenge of preserving these smoothness conditions.
In Sec. II, we analyze the linearized version of these equations and show how their solutions may be obtained locally from solutions of the scalar wave equation. This provides an important means of constructing linearized solutions in a null cone gauge for the purpose of code calibration. It also reveals essential changes in the grid structure necessary in adapting the null parallelogram algorithm for the wave equation to linearized gravity.
This also solves the major computational problems for the nonlinear case. In Sec. III, we show how the linearized algorithm can be extended to this case. In Sec. IV, we discuss the major finite difference techniques necessary for second order accuracy. In Sec. V, we present a study of the stability and accuracy of an evolution code based upon this algorithm. New exact and linearized solutions are introduced to establish second order convergence. In addition, a global check on accuracy is carried out using Bondiโs formula relating mass loss to the time integral of the square of the news function.
## II The Linearized Bondi Equations
In the linearized limit $`\beta =0`$ and $`V=r`$. The equations (3)-(8) reduce to one hypersurface equation and one evolution equation for the surviving field variables $`\gamma `$ and $`U`$,
$$(r^4U_{,r})_{,r}=\frac{2r^2(\mathrm{sin}^2\theta \gamma )_{,r\theta }}{\mathrm{sin}^2\theta }$$
(9)
$$4r(r\gamma )_{,ur}=[2r^2\gamma _{,r}r^2\mathrm{sin}\theta (\frac{U}{\mathrm{sin}\theta })_{,\theta }]_{,r}.$$
(10)
Physical considerations suggest that these equations be related to the wave equation. If this relationship were sufficiently simple, then the scalar wave algorithm could be used as a guide in formulating an algorithm for evolving $`\gamma `$. A scheme for generating solutions to the linearized Bondi equations in terms of solutions to the wave equation has been presented previously . However, in that scheme, the relationship of the scalar wave to $`\gamma `$ is non-local in the angular directions and is not useful for this purpose.
We now formulate an alternative scheme in terms of spin-weight 0 quantities $`\alpha `$ and $`Z`$, related to $`\gamma `$ (spin-weight 2) and $`U`$ (spin-weight 1) by
$$\gamma =รฐ^2\alpha =\mathrm{sin}\theta _\theta (\frac{1}{\mathrm{sin}\theta }_\theta \alpha )$$
(11)
$$U=รฐZ=_\theta Z.$$
(12)
Then the linearized equations are equivalent to
$$(r^4Z_{,r})_{,r}=2r^2(2L^2)\alpha _{,r},$$
(13)
and
$$E:=2(r\alpha )_{,ur}r^1(r^2\alpha _{,r}r^2Z/2)_{,r}=0,$$
(14)
where $`L^2=(1/\mathrm{sin}\theta )_\theta (\mathrm{sin}\theta _\theta )`$ is the $`\theta `$-part of the angular momentum operator. Now let $`\mathrm{\Phi }`$ be a solution of the flat space scalar wave equation,
$$r\mathrm{}\mathrm{\Phi }=2(r\mathrm{\Phi })_{,ur}(r\mathrm{\Phi })_{,rr}+r^1L^2\mathrm{\Phi }=0,$$
(15)
and set
$$r^2\alpha _{,r}=(r^2\mathrm{\Phi })_{,r}$$
(16)
and
$$r^2Z_{,r}=2(L^22)\mathrm{\Phi }.$$
(17)
Then
$$E=r\mathrm{}\mathrm{\Phi }+2(\mathrm{\Phi }+\alpha )_{,u}2r^2(r^2\mathrm{\Phi })_{,r}+Z,$$
(18)
and
$$E_{,r}=r^2(r^3\mathrm{}\mathrm{\Phi })_{,r}$$
(19)
Equation (13) is satisfied as a result of (16) and (17)and the wave equation (15) implies $`E_{,r}=0`$. If $`\mathrm{\Phi }`$ is smooth and $`O(r^2)`$ at the origin, this implies $`E=0`$, so that the linearized equations are satisfied. The condition that $`\mathrm{\Phi }=O(r^2)`$ eliminates fields with only monopole and dipole dependence so that it does not restrict the generality of the spin-weight 2 function $`\gamma `$ obtained through (11).
Thus for any linearized axisymmetric gravitational wave in the null cone gauge, $`\gamma `$ may be related to a scalar wave by (11) and (16). The CFL condition for convergence of a finite difference algorithm requires that the numerical domain of dependence be larger than the physical domain of dependence. Because the relationship between $`\mathrm{\Phi }`$ and $`\gamma `$ is local with respect to the null rays on the cone, their domains of dependence coincide. This suggests that a stable and convergent evolution algorithm for the gravitational field may be modeled upon the scalar wave algorithm. This turns out to be the case although some subtle distinctions arise, as described below.
An evolution algorithm for scalar waves has been formulated in terms of an integral identity for the values of the field at the corners of a null parallelogram lying on the $`(u,r)`$ plane . The wave equation with source, $`\mathrm{}\mathrm{\Phi }=S`$, can be reexpressed in the form
$$\mathrm{}^{(2)}\psi =\frac{L^2\psi }{r^2}+rS,$$
(20)
where $`\psi =r\mathrm{\Phi }`$ and $`\mathrm{}^{(2)}`$ is the 2-dimensional wave operator intrinsic to the $`(u,r)`$ plane. Integration over the null parallelogram as depicted in Fig. 1 then leads to the integral equation
$$\psi _Q=\psi _P+\psi _S\psi _R+\frac{1}{2}_A๐u๐r[\frac{L^2\psi }{r^2}+rS],$$
(21)
where $`P`$, $`Q`$, $`R`$ and $`S`$ are the corners and $`A`$ the area of the parallelogram.
This identity gives rise to an explicit marching algorithm for evolution. Let the null parallelogram span null cones at adjacent grid values $`u_0`$ and $`u_0+\mathrm{\Delta }u`$, as shown in Fig. 1, for some $`\theta `$ and $`\varphi `$. If $`\psi `$ has been determined on the entire $`u_0`$ cone and on the $`u_0+\mathrm{\Delta }u`$ cone radially outward from the origin to the point $`P`$, then (21) determines $`\psi `$ at the next point $`Q`$ in terms of an integral over $`A`$. This procedure can then be repeated to determine $`\psi `$ at the next radially outward point $`T`$ in Fig. 1. After completing this radial march to scri, the field $`\psi `$ is then evaluated on the next null cone at $`u_0+2\mathrm{\Delta }u`$, beginning at the vertex where smoothness gives the start up condition that $`\psi =0`$. The resulting evolution algorithm is a 2-level scheme which reflects, in a natural way, the distinction between characteristic and Cauchy evolution, i.e. that the time derivative of the field is not part of the characteristic initial data.
The CFL condition on the numerical domain of dependence is a necessary condition for convergence of a numerical algorithm. For the grid point at $`(u,r,\theta )`$, there are three critical grid points, $`(u\mathrm{\Delta }u,r+\mathrm{\Delta }r,\theta )`$ and $`(u\mathrm{\Delta }u,r\mathrm{\Delta }r,\theta \pm \mathrm{\Delta }\theta )`$, which must lie inside its past physical light cone. These gives rise to the inequalities $`\mathrm{\Delta }u<2\mathrm{\Delta }r`$ and $`\mathrm{\Delta }u<\mathrm{\Delta }r+(\mathrm{\Delta }r^2+r^2\mathrm{\Delta }\theta ^2)^{1/2}`$. At large $`r`$, the second inequality becomes $`\mathrm{\Delta }u<r\mathrm{\Delta }\theta `$ and the limitations on the time step are essentially the same as for a Cauchy evolution algorithm. However, near the vertex of the cone, the second inequality gives a stricter condition
$$\mathrm{\Delta }u<K\mathrm{\Delta }r\mathrm{\Delta }\theta ^2,$$
(22)
where $`K`$ is a number of order $`1`$ whose exact value depends upon the start up details at the vertex. For the scalar wave equation, these stability limits were confirmed by numerical experimentation and it was found that $`K4`$.
The linearized gravitational evolution equation (10) can be recast into a form similar to (20),
$$\mathrm{}^{(2)}\psi =,$$
(23)
where now $`\psi =r\gamma `$ and $``$ only contains hypersurface terms. This allows use of the null parallelogram algorithm to evolve $`\gamma `$ by the same marching scheme as in the scalar case. The additional feature here is that $`U`$ must be simultaneously marched out the null cone using the hypersurface equation (9). For the scalar wave equation (20), the angular momentum barrier is represented by the term $`L^2\psi /r^2`$, which is determined from $`\psi `$ by a double angular derivative. In the linearized gravitational evolution equation (10), the analogous term is $`[r^2\mathrm{sin}\theta (U/\mathrm{sin}\theta )_{,\theta }]_{,r}`$, which is determined from $`U`$ by a single angular derivative. In turn, the hypersurface equation (9) relates $`U`$ to $`\gamma `$ by a single angular derivative. Physically, this has the same net effect of producing an angular momentum barrier depending upon the second angular derivative, as in the scalar case. However, the nontrivial mathematical distinction between the two cases leads to nontrivial difference in their natural grid structures for a numerical algorithm. In particular, the grid for $`U`$ must be staggered between the grid points for $`\gamma `$. These and other details of the marching algorithm will be given in Sec. IV, where we discuss the nonlinear case.
The use of scalar waves to generate solutions of the linearized Bondi equations provides an important tool for testing evolution algorithms in the linear regime. Monopole solutions may be represented in the form $`\mathrm{\Phi }=[F(u+2r)F(u)]/r`$ and axisymmetric multipoles may be built up by applying the $`z`$-translation operator
$$_z=\mathrm{cos}\theta (_r_u)r^1\mathrm{sin}\theta _\theta $$
(24)
to these solutions. Then $`\gamma `$ may be obtained via (11). For the calibration measurements in Sec. V, we use the solutions
$$\mathrm{\Phi }=(_z)^{\mathrm{}}[u(u+2r)]^1,$$
(25)
obtained by applying $`(_z)^{\mathrm{}}`$ to the fundamental Lorentz invariant solution $`1/(x^\alpha x_\alpha )`$. This solution is well behaved above the singular light cone $`u=0`$.
## III The Nonlinear Algorithm
The nonlinear evolution equation (8) can also be recast in the form of (23) in terms of $`\psi =r\gamma `$, for an appropriate choice of 2-dimensional wave operator $`\mathrm{}^{(2)}`$. In this case, the $`(u,r)`$ submanifold is not flat so that it would not be appropriate to base $`\mathrm{}^{(2)}`$ upon a flat metric. Indeed, such a choice would lead to an improper domain of dependence and could not be used as the basis for a stable algorithm. It would seem more natural to use the $`\mathrm{}^{(2)}`$ operator of the metric induced in the $`(u,r)`$ submanifold by the 4-dimensional metric (2). Here we pursue an alternative choice based upon the line element
$$d\sigma ^2=2l_{(\mu }n_{\nu )}=e^{2\beta }du[\frac{V}{r}du+2dr],$$
(26)
where $`l_\mu =u_{,\mu }`$ is the normal to the outgoing null cones and $`n_\mu `$ is the other null vector normal to the spheres of constant $`r`$. Although this choice is not unique and other possibilities deserve exploration, it leads to the simplest $``$-terms when reexpressing (8) in the form (23). Because the domain of dependence of $`d\sigma ^2`$ contains the domain of dependence of the induced metric of the $`(u,r)`$ submanifold, this approach does not lead to convergence problems associated with the CFL condition.
The wave operator associated with (26) is
$$\mathrm{}^{(2)}\psi =e^{2\beta }[2\psi _{,ru}(\frac{V}{r}\psi _{,r})_{,r}]$$
(27)
and the nonlinear evolution equation (8) becomes
$$\mathrm{}^{(2)}\psi =e^{2\beta },$$
(28)
where
$``$ $`=`$ $`({\displaystyle \frac{V}{r}})_{,r}\gamma {\displaystyle \frac{1}{r}}[r^2(\gamma _{,\theta }U+{\displaystyle \frac{1}{2}}\mathrm{sin}\theta ({\displaystyle \frac{U}{\mathrm{sin}\theta }})_{,\theta })]_{,r}`$ (31)
$`r{\displaystyle \frac{(\gamma _{,r}U\mathrm{sin}\theta )_{,\theta }}{\mathrm{sin}\theta }}+{\displaystyle \frac{1}{4}}r^3e^{2(\gamma \beta )}(U_{,r})^2`$
$`+{\displaystyle \frac{1}{r}}e^{2(\beta \gamma )}[(\beta _{,\theta })^2+\mathrm{sin}\theta ({\displaystyle \frac{\beta _{,\theta }}{\mathrm{sin}\theta }})_{,\theta }].`$
Because all 2-dimensional wave operators are conformally flat, with conformal weight $`2`$, the surface integral of (28) over a null parallelogram gives an integral equation analogous to (21),
$$\psi _Q=\psi _P+\psi _S\psi _R+\frac{1}{2}_A๐u๐r.$$
(32)
This allows the evolution of $`\gamma `$ by the same basic marching algorithm as described in Sec. II for the scalar wave and linearized wave cases. The additional modifications are that $`\beta `$, $`U`$, and $`V`$ must be simultaneously marched out the null cone using the nonlinear hypersurface equations (3)-(6). Because of the hierarchal structure of these equations, $`\gamma `$, $`\beta `$, $`U`$, and $`V`$ may be marched in that order without any matrix inversions or other implicit operations. The basic computational cell and finite difference constructions are described in the next section.
Near the origin, the metric approaches the Minkowski metric so that the stability of the nonlinear algorithm in this region should be subject to the same Courant limit as for the linearized equations. Near scri, the gravitational variables have the asymptotic form
$`\gamma `$ $`=`$ $`K+r^1c+O(r^2)`$ (33)
$`\beta `$ $`=`$ $`H+O(r^2)`$ (34)
$`U`$ $`=`$ $`L+O(r^1)`$ (35)
$`V`$ $`=`$ $`r^2{\displaystyle \frac{(L\mathrm{sin}\theta )_{,\theta }}{\mathrm{sin}\theta }}`$ (36)
$`+`$ $`re^{2(HK)}\left[1+2{\displaystyle \frac{(H_{,\theta }\mathrm{sin}\theta )_{,\theta }}{\mathrm{sin}\theta }}+{\displaystyle \frac{(K_{,\theta }\mathrm{sin}^3\theta )_{,\theta }}{\mathrm{sin}^3\theta }}+4(H_{,\theta })^24H_{,\theta }K_{,\theta }2(K_{,\theta })^2\right]`$ (37)
$``$ $`2e^{2H}+O(r^1),`$ (38)
where $``$ corresponds to Bondiโs mass aspect. In a standard Bondi frame at scri $`K`$, $`H`$, and $`L`$ all vanish, but not in null cone coordinates adapted to a Minkowski frame at the origin. This dependence can lead to drastic behavior of the $`u`$-coordinate at large distances. In a numerical study of spherically symmetric, self-gravitating scalar radiation fields , $`H\mathrm{}`$ as a horizon is formed and an infinite redshift develops between central observers and observers at scri. In that case, a consideration of the domains of dependence indicates that the step size $`\mathrm{\Delta }u`$ for stable evolution must approach zero as the horizon is formed. The divergence of the outgoing null cone equals $`e^{2\beta }/r`$. If $`\beta \mathrm{}`$ at a finite value of $`r`$ then a caustic will in general form. When this occurs, a single null cone coordinate system cannot cover the entire exterior region of the spacetime.
In the more general case being considered here, it is also possible for the $`u`$-direction to become spacelike at large distances, corresponding to the coefficient of $`du^2`$ in (2) becoming negative. (The $`u=constant`$ hypersurfaces, of course, must remain null.) As discussed in Sec. V, this does not affect the stability of the algorithm. The algorithm is also valid when the vertex of the null cone is replaced by an inner boundary consisting of a timelike or null worldtube, so that it may also be applied to other versions of the characteristic initial value problem .
## IV Finite Difference Techniques
The numerical grid is based upon the outgoing null cones, using the compactified radial coordinate $`x=r/(1+r)`$ and the angular coordinate $`y=\mathrm{cos}\theta `$. Thus points at scri are included in the grid at $`x=1`$.
In order to improve numerical accuracy at the grid boundaries, the code is written in terms of the variables $`\widehat{\psi }=r\widehat{\gamma }=r\gamma /\mathrm{sin}^2\theta `$, $`\beta `$, $`S=(Vr)/r^2`$ and $`\widehat{U}=U/\mathrm{sin}\theta `$. For a pure quadrupole mode, $`\widehat{\gamma }`$ has constant angular dependence.
To develop a discrete evolution algorithm, we work with two sets of spatial grid points, both of which have the constant spacing $`\mathrm{\Delta }x=1/N_x`$ and $`\mathrm{\Delta }y=1/N_y`$. The first grid is defined by $`(u^n,x_i,y_j)=(n\mathrm{\Delta }u,i\mathrm{\Delta }x,j\mathrm{\Delta }y)`$. The second grid is shifted (staggered) in both the $`x`$ and $`y`$ variables and is thus defined by $`(u^n,x_{i+\frac{1}{2}},y_{j+\frac{1}{2}})=(n\mathrm{\Delta }u,(i+\frac{1}{2})\mathrm{\Delta }x,(j+\frac{1}{2})\mathrm{\Delta }y)`$. Note that the staggered grid extends beyond the physical boundary $`x=1`$. This peculiarity is successfully exploited for a smooth calculation of the metric at scri. The time step is variable and is limited by the largest possible value that would satisfy the CFL condition over the entire grid.
A staggerered grid is not necessary for the scalar wave equation but its introduction for the gravitational case is dictated by a detailed von Neumann stability analysis of the linearized equations. Accordingly, $`\widehat{\gamma },\beta \text{and}S`$ reside on the $`(x_i,y_j)`$ grid while $`\widehat{U}`$ is placed on the $`(x_{i+\frac{1}{2}},y_{j+\frac{1}{2}})`$ grid. We denote values of a field $`F`$ at the site $`(n,i,j)`$ by $`F_{ij}^n=F(u^n,x_i,y_j)`$. We use centered second order differences for derivatives at points not on the edges of the grid, e.g. for an arbitrary field $`F`$
$$F_{,y|}^{}{}_{i,j}{}^{}=\frac{1}{2\mathrm{\Delta }y}(F_{i,j+1}F_{i,j1})$$
(39)
To calculate derivatives of the field at the edges of the grid, we use backward and forward second-order differences, e.g at the $`y=\pm 1`$ edges of the grid, where $`j=\pm N_y`$
$$F_{,y|}^{}{}_{i,\pm N_y}{}^{}=\frac{1}{2\mathrm{\Delta }y}(F_{i,\pm N_y2}+4F_{i,\pm N_y1}3F_{i,\pm N_y})$$
(40)
### A The Hypersurface Equations
In terms of the numerical variables $`\widehat{\gamma }`$, $`\beta `$, $`S`$ and $`\widehat{U}`$, and expressed in the coordinates $`x`$ and $`y`$ used in the code, the hypersurface equations (3)-(6) are
$`\beta _{,x}`$ $`=`$ $`_\beta `$ (41)
$`({\displaystyle \frac{x^4}{(1x)^2}}e^{2((1y^2)\widehat{\gamma }\beta )}\widehat{U}_{,x})_{,x}`$ $`=`$ $`2{\displaystyle \frac{x^2}{(1x)^2}}_U`$ (42)
$`x^2S_{,x}+{\displaystyle \frac{2x}{1x}}S`$ $`=`$ $`_S,`$ (43)
where the source terms $`_\beta `$, $`_U`$ and $`_S`$ are given by
$`_\beta `$ $`=`$ $`{\displaystyle \frac{1}{2}}x(1x)(1y^2)^2(\widehat{\gamma }_{,x})^2`$ (44)
$`_U`$ $`=`$ $`\beta _{,xy}{\displaystyle \frac{2}{x(1x)}}\beta _{,y}+4y\widehat{\gamma }_{,x}+(1y^2)[2\widehat{\gamma }_{,x}((1y^2)\widehat{\gamma }_{,y}2y\widehat{\gamma })\widehat{\gamma }_{,xy}]`$ (45)
$`_S`$ $`=`$ $`1xy[{\displaystyle \frac{4}{1x}}\widehat{U}+x\widehat{U}_{,x}]+{\displaystyle \frac{1}{2}}(1y^2)x[{\displaystyle \frac{4}{1x}}\widehat{U}_{,y}+x\widehat{U}_{,xy}]`$ (49)
$`{\displaystyle \frac{1}{4}}x^4(\widehat{U}_{,x})^2(1y^2)e^{2((1y^2)\widehat{\gamma })\beta }e^{2(\beta (1y^2)\widehat{\gamma })}\{112\widehat{\gamma }2y\beta _{,y}`$
$`+(1y^2)[10\widehat{\gamma }+8y\widehat{\gamma }_{,y}+8\widehat{\gamma }^2+4y\widehat{\gamma }\beta _{,y}+\beta _{,yy}+(\beta _{,y})^2`$
$`(1y^2)^2(8\widehat{\gamma }^2+2\widehat{\gamma }_{,y}\beta _{,y}+\widehat{\gamma }_{,yy}+8y\widehat{\gamma }\widehat{\gamma }_{,y})+2(1y^2)^3(\widehat{\gamma }_{,y})^2\}.`$
Note that Eq. (41) has just one radial derivative, and can be evaluated at the points $`(n,i\frac{1}{2},j)`$. We can discretize it as follows
$$\beta _{i,j}=\beta _{i1,j}+(_\beta )_{i\frac{1}{2},j}\mathrm{\Delta }x.$$
(50)
Equation (42), which contains a second radial derivative, is evaluated at the points $`(n,i1,j)`$. Near the origin, where $`\widehat{U}=O(x)`$, it becomes a singular differential equation. In order to integrate it to second order accuracy it is necessary to apply numerical regularization to the derivatives on the left hand side. Noting that $`/x=4x^3/x^4`$, Eq. (42) becomes
$`2x(1x)[x^4\widehat{U}_x]_{,x^4}`$ $`+`$ $`x^2[1(1x)(\beta _{,x}(1y^2)\widehat{\gamma }_{,x})]\widehat{U}_{,x}`$ (51)
$`=`$ $`e^{2(\beta (1y^2)\widehat{\gamma })}(1x)_U.`$ (52)
Using the identity
$$x_{i\frac{1}{2}}^4x_{i1\frac{1}{2}}^4=2\mathrm{\Delta }xx_{i1}(x_{i\frac{1}{2}}^2+x_{i1\frac{1}{2}}^2),$$
(53)
we can discretize (52) as
$`{\displaystyle \frac{(1x_{i1})}{(x_{i\frac{1}{2}}^2+x_{i1\frac{1}{2}}^2)}}[x_{i\frac{1}{2}}^4(\widehat{U}_{i,j}\widehat{U}_{i1,j})x_{i1\frac{1}{2}}^4(\widehat{U}_{i1,j}\widehat{U}_{i2,j})]`$ (54)
$`+`$ $`{\displaystyle \frac{1}{2}}x_{i1}^2\mathrm{\Delta }x[1(1x_{i1})(\beta _{i,j}\beta _{i2,j}(1y_j^2)(\widehat{\gamma }_{i,j}\widehat{\gamma }_{i2,j})){\displaystyle \frac{1}{\mathrm{\Delta }x}}](\widehat{U}_{i,j}\widehat{U}_{i2,j})`$ (55)
$`=`$ $`(\mathrm{\Delta }x)^2(1x_{i1})(_U)_{i1,j}.`$ (56)
The above is a 3-point formula, and it can not be applied at the points at $`x_i=\mathrm{\Delta }x`$, however we know the asymptotic behavior of $`U`$ at the origin, and we can use it to construct a starting algorithm for these points. We approximate the Bondi equations for $`\widehat{\gamma }`$ and $`\widehat{U}`$ by the leading two terms in a power series,
$`\widehat{\gamma }`$ $`=`$ $`ar^2+br^3`$ (57)
$`\widehat{U}`$ $`=`$ $`4y(ar+{\displaystyle \frac{1}{3}}br^2),`$ (58)
Expanding $`\widehat{\gamma }_{,u}`$ to the same order, we obtain for the rate of change of $`a`$ and $`b`$
$`a_{,u}`$ $`=`$ $`{\displaystyle \frac{6}{5}}b`$ (59)
$`b_{,u}`$ $`=`$ $`0.`$ (60)
By fitting a least square polynomial to $`\widehat{\gamma }/r^2`$ near the origin, we can read off the coefficients $`a`$ and $`b`$ and evaluate $`\widehat{U}`$ on the next hypersurface. This approximation is consistent with the global second order accuracy of the algorithm.
As with Eq. (41), Eq. (43) can be approximated at sites $`(n,i\frac{1}{2},j)`$ as follows
$`x_{i\frac{1}{2}}^2{\displaystyle \frac{1}{\mathrm{\Delta }x}}(S_{i,j}S_{i1,j})`$ $`+`$ $`{\displaystyle \frac{x_{i\frac{1}{2}}}{1x_{i\frac{1}{2}}}}(S_{i,j}+S_{i1,j})`$ (61)
$`=`$ $`(_S)_{i\frac{1}{2},j}.`$ (62)
### B The Evolution Equation
In practice, the corners of the null parallelogram, $`P`$, $`Q`$, $`R`$ and $`S`$, cannot be chosen to lie exactly on the grid because the velocity of light in terms of the compactified coordinate $`x`$ is not constant even in flat space. Numerical experimentation suggests that a stable algorithm with high accuracy results from the choice made in Fig. 1. The essential feature of this placement of the parallelogram with respect to a coordinate cell is that the sides formed by incoming rays intersect adjacent u-hypersurfaces at equal but opposite x-displacements from the neighboring grid points. Solution for the null geodesics of the metric (26) to second order accuracy then gives for the coordinates of the vertices
$`x_{i1}x_P`$ $`=`$ $`x_Rx_{i1}`$ (63)
$`=`$ $`\mathrm{\Delta }u(1x_{i1})^2[1+r_{i1}(S_{i2}^{n+1}+S_i^n)/2]/4`$ (64)
$`x_ix_Q`$ $`=`$ $`x_Sx_i`$ (65)
$`=`$ $`\mathrm{\Delta }u(1x_i)^2[1+r_i(S_{i1}^{n+1}+S_{i+1}^n)/2]/4.`$ (66)
The elementary computational cell consists of the lattice points $`(n,i,j)`$ and $`(n,i\pm 1,j)`$ on the โoldโ hypersurface and the points $`(n+1,i2,j)`$, $`(n+1,i1,j)`$ and $`(n+1,i,j)`$ on the โnewโ hypersurface (and their nearest neighbors in the angular direction). The marching algorithm computes the value of the fields at the point $`(n+1,i,j)`$ in terms of their predetermined values at the other points in the cell.
The values of $`\widehat{\psi }`$ at the vertices of the parallelogram are approximated to second order accuracy by linear interpolation between grid points. Furthermore, cancellations arise between these four interpolations so that the evolution equation (32) is satisfied to fourth order accuracy, provided the integral can be calculated to that accuracy. This is accomplished by approximating the integrand by its value at the center of the parallelogram. To second order accuracy, this gives
$$_A๐u๐r=_c_A๐u๐r=\frac{1}{2}\mathrm{\Delta }u(r_Qr_P+r_Sr_R)_c,$$
(67)
where the centered value $`_c`$ can be obtained by averaging between appropriate points in the cell. Thus the discretized version of (32) is given by
$`\widehat{\psi }_i^{n+1}`$ $`=`$ $`(\widehat{\psi }_{i1}^{n+1},\widehat{\psi }_{i2}^{n+1},\widehat{\psi }_{i+1}^n,\widehat{\psi }_i^n,\widehat{\psi }_{i1}^n)`$ (69)
$`+{\displaystyle \frac{1}{2}}\mathrm{\Delta }u(r_Qr_P+r_Sr_R)_c`$
where $``$ is a linear function of $`\psi `$ and the $`j`$ index has been suppressed.
Consequently, it is possible to move through the interior of the lattice, computing $`\widehat{\psi }_i^{n+1}`$ explicitly by an orderly radial march. This is achieved by starting at the origin at time $`u^{n+1}`$. Field values vanish there. Next, proceed outward one radial step using the boundary conditions (discussed below). Then step outward to the next interior radial point using (69), iterating this process throughout the interior and for all angles. This updates all field values stretching to scri along the new null cone at $`u^{n+1}`$, thus completing one evolutionary time step.
The above scheme is sufficient for accurate evolution in a neighborhood of the origin. Global evolution, including the points at $`x=1`$, requires careful manipulation of (69) to avoid problems from the fact that $`\psi rK`$ at scri. Thus the direct use of this formula is not possible for the point at scri, while points near scri would suffer serious loss of accuracy. We renormalize (69) in the following way. First, we introduce the quantity $`\varphi =\psi (1x)`$. Near scri $`\varphi `$ has the desired finite behavior, while near the origin it leaves unchanged the constant coefficient form of the evolution equation, thus preserving the stability properties. With this substitution and with the use of (67), the evolution equation (32) becomes
$`\varphi _Q`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{\Delta }ux_Q_c+{\displaystyle \frac{(1x_Q)}{(1x_P)}}(\varphi _P{\displaystyle \frac{1}{4}}\mathrm{\Delta }ux_P_c)`$ (71)
$`+{\displaystyle \frac{(1x_Q)}{(1x_S)}}(\varphi _S+{\displaystyle \frac{1}{4}}\mathrm{\Delta }ux_S_c){\displaystyle \frac{(1x_Q)}{(1x_R)}}(\psi _R+{\displaystyle \frac{1}{4}}\mathrm{\Delta }ux_R_c)`$
Now all terms have finite asymptotic value. The coefficient $`(1x_Q)/(1x_S)`$ has $`0/0`$ behavior at scri but approaches the limit $`1`$. Further refinement is possible with the use of the explicit second order approximation for the characteristics (66) which leads to the approximation
$$\frac{(1x_Q)}{(1x_S)}=1+2\frac{\delta }{1\delta }$$
(72)
where
$$\delta =\frac{1}{4}\mathrm{\Delta }u(1+x_i(S_{i1}^{n+1}+S_{i+1}^n)/2)$$
(73)
The final result is that the equation (71) propagates $`\varphi `$ radially outward one cell with an error of fourth order in grid size. This is valid for all interior points and the point at scri. The error in each cell compounds to a third order error on each null cone and a second order global error after evolving for a given physical time. Second order global accuracy is indeed confirmed by the convergence tests described in Sec. V.
We mentioned that a modified form of the basic grid cell Eq. (66) is used at the origin. This is necessary since the incoming characteristic through the points $`P`$ and $`R`$ can not be centered at $`x=0`$. The corners of the modified cell are given by
$`x_P`$ $`=`$ $`0`$ (74)
$`x_R`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Delta }u`$ (75)
$`x_1x_Q`$ $`=`$ $`x_Sx_1={\displaystyle \frac{1}{4}}\mathrm{\Delta }u(1x_i)^2.`$ (76)
Only the linear terms of $``$ are kept while evolving the first point, i.e. for $`x_1=\mathrm{\Delta }x`$. This reduces equation (32) to
$$\widehat{\psi }_Q=\widehat{\psi }_S\widehat{\psi }_R\frac{1}{4}_A๐u๐r\frac{1}{r}(r^2\widehat{U}_{,y})_{,r}.$$
(77)
Using the expansion (58) for $`U`$ near the origin, the integral simplifies further to
$$_A๐u๐r(3ar+\frac{12}{5}br^2).$$
(78)
The integrand is now evaluated to second order accuracy at $`u^{n+\frac{1}{2}}=u^n+\frac{\mathrm{\Delta }u}{2}`$ using
$$a^{n+\frac{1}{2}}=a^n+\frac{6}{5}b^n\frac{\mathrm{\Delta }u}{2}$$
and $`b^{n+\frac{1}{2}}=b^n`$. Keeping higher order terms would not improve the global convergence rate of the code.
## V Code tests
### A Testbeds
As we have shown in Sec. II, linearized solutions of the Bondi equations can be generated by solutions of the scalar wave equation, thus supplying a complete set of test beds for the very weak regime. For the nonlinear case, exact boost and rotation symmetric solutions of the Bondi initial hypersurface equations have also been found . They have been used to check the radial integrations leading from $`\gamma `$ to $`\beta `$, $`U`$ and $`V`$ but they do not provide a test of the evolution algorithm. However, in the course of this work, we have found that one of these initial data sets is in fact preserved under time evolution and is an exact static solution of the nonlinear vacuum equations.
This solution provides an important test bed for null cone evolution codes. Except for spherically symmetric cases, it is the only known solution of Einsteinโs equation which can be expressed explicitly in null cone coordinates with no singularity at the vertex. It has the form
$`2e^\gamma `$ $`=`$ $`1+\mathrm{\Sigma }`$ (79)
$`e^{2\beta }`$ $`=`$ $`{\displaystyle \frac{(1+\mathrm{\Sigma })^2}{4\mathrm{\Sigma }}}`$ (80)
$`U`$ $`=`$ $`{\displaystyle \frac{a^2\rho \sqrt{r^2\rho ^2}}{r\mathrm{\Sigma }}}`$ (81)
$`V`$ $`=`$ $`{\displaystyle \frac{r\left(2a^2\rho ^2a^2r^2+1\right)}{\mathrm{\Sigma }}}`$ (82)
where $`\rho =r\mathrm{sin}\theta `$, $`\mathrm{\Sigma }=\sqrt{1+a^2\rho ^2}`$ and $`a`$ is a free scale parameter. It is remarkable that the null data $`\gamma `$ is time independent under evolution, as can be verified by inserting the above expressions in (8). Because this solution is static, as well as boost and rotation symmetric, the commutator between the boost and time translation symmetries implies that it has an additional translation symmetry in the boost direction. Thus the solution falls into several overlapping and widely studied classes, including the static cylindrically symmetric spacetimes and the Weyl static axisymmetric spacetimes. However, this solution, which we call SIMPLE, has not previously been singled out, apparently because it cannot easily be identified in the traditional coordinates used for studying static solutions. Because of its cylindrical symmetry, it is clear that this solution is not asymptotically flat but it can be used to construct an asymptotically flat, nonsingular solution by smoothly pasting asymptotically flat null data to it outside some radius $`R`$. The resulting solution will be static and given by (82) in the domain of dependence interior to $`R`$. Numerical solutions generated by this technique are used in the code calibration tests presented below.
In addition, global energy conservation provides an important test bed. The Bondi mass loss formula is not one of the equations used in the evolution algorithm but follows from those equations as a consequence of a global integration of the Bianchi identities.
### B Convergence and Stability
We have tested the algorithm to be second order accurate and stable, subject to the CFL condition, throughout the regime in which caustics and horizons do not form. In Sec. II, we showed how the the linearized Bondi equations may be reduced to the scalar wave equation by local operations. For very weak data, the nonlinear equations approximate the linear equations so that we would expect the global stability of the nonlinear algorithm to be related to the CFL condition for the scalar wave algorithm. Near the origin, stability checks show that the time step is limited by (22) with $`K=8`$, which is twice the limit found for the scalar wave algorithm. This factor of two apparently arises from the use of a staggered grid in the gravitational case, which effectively doubles the value of $`r`$ at which the main algorithm takes over from the start up algorithm at the origin. This gives some reassurance that the scalar wave algorithm has been optimally adapted to the Bondi equations.
By construction, the $`u`$-direction is timelike at the origin where it coincides with the worldline traced out by the vertex of the outgoing null cone. But even for weak fields, the $`u`$-direction becomes spacelike at large distances along a typical outgoing ray. This can be seen from the metric coefficient $`g_{uu}=(V/r)e^{2\beta }U^2r^2e^{2\gamma }`$ which at large $`r`$ becomes dominated by the the asymptotic behavior $`U=L+O(1/r)`$. Geometrically, this reflects the property that scri is itself a null hypersurface so that all internal directions are spacelike, except for the null generator. For a flat space time, the $`u`$-direction picked out at the origin corresponds to the null direction at scri but it becomes asymptotically spacelike under the slightest deviation from spherical symmetry.
By choosing initial data of very small amplitude ($`|\widehat{\gamma }|10^9`$), we have performed convergence tests of the numerical solutions against the solutions of the linearized equations. The linearized solutions (25) were given as initial data at $`u=0`$ and we compared the numerically evolved solutions to the linearized solutions at a central time of $`u=0.5`$. We observed that for the low angular momentum solutions ($`\mathrm{}=`$2, 3, 4) the code is superaccurate, i.e. the solutions converge to the exact result at a rate faster than second order in the grid size. This is to be expected since for these solutions the hatted variables used in the code exhibit angular dependence that is at most quadratic in $`y`$, so that the second order accurate $`y`$-derivatives are calculated exactly. The error, as measured by the $`L_2`$ norm, of a numerical solution with higher harmonics ($`\mathrm{}=6`$) is graphed in Fig. 2. The slope of the graph gives a convergence rate of $`2.04\pm 0.01`$ with respect to grid size. This result is insensitive to the particular norm used, i.e. we also verified second order convergence in the $`L_1`$ and $`L_{\mathrm{}}`$ norms.
Second order convergence has also been checked against the exact static solution SIMPLE. Since this solution is not asymptotically flat, we match the initial data smoothly to asymptotically flat data for a nonstatic exterior. Consideration of the domain of dependence implies that the matching boundary propagates along an ingoing null hypersurface. Thus, we can obtain a reliable measure of how accurately the evolution preserves the static interior, provided we restrict the calculation of the error norm to a region not yet influenced by the exterior nonstatic data. We matched one such static solution in the interior ($`x0.5`$) to smooth exterior data with compact support. We calculated the $`L_{\mathrm{}}`$ norm for the region $`x0.4`$ at time u=0.25, and considered the dependence of the error on the grid size. The preservation of the static interior to graphical accuracy is shown in Fig. 3, while the second order convergence of the error is demonstrated in Fig. 4. In addition, for a wide variety of initial data having unknown analytic solution, we have verified that the numerical solution converges to second order in the sense of Cauchy convergence.
Stability of the code in the low to medium amplitude regime has been verified experimentally by running arbitrary initial data until it radiates away to scri. At higher amplitudes, it is expected that physical singularities will arise, but we have not yet explored this regime. Figure 5 shows a sequence of time slices of the numerical evolution of some arbitrary initial data of compact support. Note the rich angular structure that arises at $`u0.25`$ and then dissipates. At $`u1.5`$ the amplitude of the field is sufficiently small so that it appears to be zero in the figure. It continues to decay at later times.
### C Energy Conservation
The Bondi mass loss formula relates the gravitational radiation power to the square of the news function. It follows from the equations used in the algorithm as a consequence of a global integration of the Bianchi identities. Thus it not only furnishes a valuable tool for physical interpretation but it also provides a very important calibration of numerical accuracy and consistency.
Historically, numerical calculations of the Bondi mass $`M_B`$ have been frustrated by technical difficulties arising from the necessity to pick off nonleading terms in an asymptotic expansion about infinity. For example, the mass aspect $``$ must be picked off in the the asymptotic expansion (38) for $`V`$. This is similar to the experimental task of determining the mass of an object by measuring its far field. In the non-radiative case it can be accomplished by measuring gravity gradients, but otherwise this approach can be swamped by radiation fields. In the computational problem, further complications arise from gauge terms which dominate asymptotically even over the radiation terms. We have recently developed a second order accurate algorithm for calculating the Bondi mass . It avoids the above problems through the use of Penrose compactification and the introduction of renormalized variables in which Bondiโs mass aspect appears as the leading asymptotic term. The Bondi mass algorithm depends only upon fields on a single null hypersurface. It has been incorporated into the present evolution code to calculate the mass at any given retarded time.
In the present formalism, the news function $`N`$ is given by
$$2e^{2H}N=2c_{,u}+\frac{(\mathrm{sin}\theta c^2L)_{,\theta }}{\mathrm{sin}\theta c}+e^{2K}\omega \mathrm{sin}\theta \left[\frac{(e^{2H}\omega )_{,\theta }}{\mathrm{sin}\theta \omega ^2}\right]_{,\theta }.$$
(83)
Here $`\omega `$ is the conformal factor relating the asymptotic 2-geometry to the unit sphere geometry of a Bondi frame, i.e.
$$e^{2K}d\theta ^2+\mathrm{sin}^2\theta e^{2K}d\varphi ^2=\omega ^2(d\theta _B^2+\mathrm{sin}^2\theta _Bd\varphi _B^2),$$
(84)
where $`\theta _B`$ and $`\varphi _B=\varphi `$ are Bondi spherical coordinates. Calculation of $`\omega `$ complicates the calculation of the news function. The simplest approach is to set $`y=\mathrm{cos}\theta `$ and $`y_B=\mathrm{cos}\theta _B`$. Then
$$\omega ^2=\frac{dy_B}{dy},$$
(85)
where
$$y_B=\mathrm{tanh}\left[_0^y\frac{dy}{1y^2}e^{2K}\right].$$
(86)
This gives
$$\omega =\frac{2e^K}{(1+y)e^\mathrm{\Delta }+(1y)e^\mathrm{\Delta }}$$
(87)
where
$$\mathrm{\Delta }=_0^y๐y\frac{e^{2K}1}{1y^2}.$$
(88)
In order to prepare this integral in an explicitly regular form for computation we introduce an auxiliary parameter $`\alpha `$ and rewrite (88) as the double integral
$$\mathrm{\Delta }=2_0^y๐y_0^1๐\alpha e^{2\alpha K}\widehat{K},$$
(89)
where $`\widehat{K}=K/(1y^2)`$ is regular at the poles. It is then straightforward to obtain a second order accurate finite difference formula for the news function.
The Bondi formula for energy conservation between central times $`u_0`$ and $`u`$ takes the form $`C`$ =0, where
$$C=M_B(u)M_B(u_0)+\frac{1}{2}_1^1๐y_{u_0}^u๐ue^{2H}\omega ^1N^2.$$
(90)
Figure 6 graphs $`C`$ relative to the initial mass $`M_B(u_0)`$ for a numerical evolution of the polynomial data
$$\widehat{\gamma }=\lambda \frac{\left[(xx_1)(xx_2)(y^2y_0^2)\right]^6}{\left[(x_1x_2)y_0\right]^{12}}$$
(91)
with compact support in the domain $`(x_1xx_2)\times (y_0yy_0)`$ and amplitude parameter $`\lambda `$. For the graph, we have chosen $`\lambda =0.3`$, $`x_1=0.1`$, $`x_2=0.5`$ and $`y_0=0.5`$, and evolved the numerical solution up to $`u=0.01`$, on a grid of 512 radial $`\times `$ 128 angular points. The bulk of the error occurs in the calculation of the Bondi mass, whose accuracy is more sensitive to grid size than the accuracy of either the news function or the evolution code. Most of this error may be removed by using Richardson extrapolation to take advantage of the known second order accuracy of the Bondi mass. For example, if $`F_n(x_i)`$ is a second order accurate finite difference approximation to the function $`f(x)`$ on a grid of $`n`$ points, then $`(4F_nF_{n/2})/3`$ approximates $`f`$ to third order and in fact to fourth order if odd orders are absent in the approximation. This absence of odd orders indeed holds for the Bondi mass because all derivatives, interpolations and integrals are centered. Thus, introduction of subgrids obtained by subsampling, leads to a fourth order expression for $`M_B`$, with the corresponding relative error in energy conservation also graphed in Fig. 6. In this way, energy conservation is attained to 0.4% accuracy. Note that only a single evolution on a fixed grid is necessary here because Richardson extrapolation is applied when calculating the Bondi mass. For the purposes of the graph, we have done this for each time that $`C`$ is plotted, but to check energy conservation, it suffices to do it only at the initial and final times. Figure 6 serves as a rewarding testament to the virtues of a code with known convergence rates.
## VI Conclusion
We have constructed a second order accurate evolution algorithm for the null cone initial value problem for axisymmetric vacuum spacetimes. Energy conservation is maintained to second order accuracy. Extensive tests of the algorithm establish that it is globally valid in the regime where horizons and caustics do not develop. This generates a large complement of highly accurate numerical solutions for the class of asymptotically flat, axisymmetric vacuum spacetimes, for which no analytic solutions are known. All results of numerical evolutions in this regime are consistent with the theorem of Christodoulou and Klainerman that weak initial data evolve asymptotically to Minkowski space at late time. The code is now being tested in the strong field regime for application to the study of black hole formation.
###### Acknowledgements.
We benefited from research support from the National Science Foundation under NSF Grant PHY92-08349 and from computer time made available by the Pittsburgh Supercomputing Center.
##
We sketch here the von Neumann stability analysis of the algorithm for the linearized Bondi equations. The analysis is based up freezing the explicit functions of $`r`$ and $`y`$ that appear in the equations, so that it is only valid locally for grid sizes satisfying $`\mathrm{\Delta }r<<r`$ and $`\mathrm{\Delta }y<<1`$. However, as is usually the case, the results are quite indicative of the stability of the actual global behavior of the code.
Starting with the hatted code variables introduced in Sec. IV and setting $`\mathrm{\Gamma }=r^2\widehat{U}`$ and $`G=r\widehat{\gamma }`$, the linearized Bondi equations (9) and (10) take the form
$$r^2\mathrm{\Gamma }_{,rr}2\mathrm{\Gamma }=2[4y(1y^2)_y](rG_{,r}G)$$
(92)
and
$$2G_{,ur}G_{,rr}=(1/2r)\mathrm{\Gamma }_{,ry}.$$
(93)
Freezing the explicit factors of $`r`$ and $`y`$ at $`r=R`$ and $`y=Y`$, introducing the Fourier modes $`G=e^{su}e^{ikr}e^{ily}`$ (with real $`k`$ and $`l`$) and setting $`\mathrm{\Gamma }=AG`$, these equations imply
$$A=2(1ikR)[4Y(1Y^2)il]/(2+R^2k^2)$$
(94)
and
$$4is=2k+Al/R.$$
(95)
For stable modes, $`Re(s)0`$. This requires that the $`lIm(A)0`$ which will be satisfied unless $`Ykl<0`$. In the latter case, unstable solutions exist to the PDEโs obtained by freezing the coefficients in the linearized equations (92) and (93). The linearized equations themselves do not have unstable modes but they arise in the frozen coefficient formalism from dropping the boundary condition of spherical topology on the $`y`$-dependence. For a global solution, $`G`$ should not have periodic dependence on $`y`$ but instead be decomposed into spin-weight 2 harmonics, in which case instabilities would not arise in the above analysis. Thus these unstable modes of the frozen PDE are artificial and should be discarded by requiring $`Ykl0`$ when analyzing the stability of the corresponding FDE.
Consider now the FDE obtained by putting $`G`$ on the grid points $`r_I`$ and $`\mathrm{\Gamma }`$ on the staggered points $`r_{I+1/2}`$, while using the same angular grid $`y_J`$ and time grid $`u_N`$. Let $`P`$, $`Q`$, $`R`$ and $`S`$ be the corner points of the null parallelogram algorithm, placed so that $`P`$ and $`Q`$ are at level $`N+1`$, $`R`$ and $`S`$ are at level $`N`$, and so that the line $`PR`$ is centered about $`r_I`$ and $`QS`$ is centered about $`r_{I+1}`$. For simplicity, we display the analysis at the equator where $`Y=0`$. Then, using linear interpolation and centered derivatives and integrals, the null parallelogram algorithm for the frozen version of the linearized equations leads to the FDEโs
$`(R/\mathrm{\Delta }r)^2(\mathrm{\Gamma }_{I+3/2}2\mathrm{\Gamma }_{I+1/2}+\mathrm{\Gamma }_{I1/2})(\mathrm{\Gamma }_{I+3/2}+\mathrm{\Gamma }_{I1/2})`$ (96)
$`=\delta _y[2(R/\mathrm{\Delta }r)(G_{I+1}G_I)(G_{I+1}+G_I)]`$ (97)
(all at the same time level) and
$`G_{I+1}^{N+1}G_I^{N+1}G_{I+1}^N+G_I^N`$ (98)
$`+`$ $`(\mathrm{\Delta }u/4\mathrm{\Delta }r)(G_{I+1}^{N+1}+2G_I^{N+1}G_{I1}^{N+1}G_{I+2}^N+2G_{I+1}^NG_I^N)`$ (99)
$`=`$ $`(\mathrm{\Delta }u/8R)\delta _y(\mathrm{\Gamma }_{I+1/2}^{N+1}\mathrm{\Gamma }_{I1/2}^{N+1}+\mathrm{\Gamma }_{I+3/2}^N\mathrm{\Gamma }_{I+1/2}^N),`$ (100)
where $`\delta _y`$ represents a centered first derivative. Again setting $`\mathrm{\Gamma }=AG`$ and introducing the discretized Fourier mode $`G=e^{sN\mathrm{\Delta }u}e^{ikI\mathrm{\Delta }r}e^{ilJ\mathrm{\Delta }y}`$, we have $`\delta _y=i\mathrm{sin}(l\mathrm{\Delta }y)/\mathrm{\Delta }y`$ and (97) and (100) reduce to
$$A[(R/\mathrm{\Delta }r)^2(1\mathrm{cos}\alpha )+\mathrm{cos}\alpha ]=L[(2R/\mathrm{\Delta }r)\mathrm{sin}(\alpha /2)+i\mathrm{cos}(\alpha /2)]$$
(101)
and
$$e^{s\mathrm{\Delta }u}=e^{i\alpha }(C^{}AD)/(CAD),$$
(102)
where $`L=\mathrm{sin}(l\mathrm{\Delta }y)/\mathrm{\Delta }y`$, $`\alpha =k\mathrm{\Delta }r`$, $`C=ie^{i\alpha /2}\mathrm{sin}(\alpha /2)+(\mathrm{\Delta }u/4\mathrm{\Delta }r)(1\mathrm{cos}\alpha )`$ and $`D=(L\mathrm{\Delta }u/8R)\mathrm{sin}(\alpha /2)`$. The stability condition that $`Re(s)0`$ then reduces to $`Re[CD(AA^{})]0`$ which is equivalent to $`1+\mathrm{cos}\alpha [1(\mathrm{\Delta }r/R)^2]0`$. Thus this stability condition is automatically satisfied and poses no constraint on the algorithm.
The corresponding analysis at the poles $`Y=\pm 1`$ again leads to (102), where now
$$A[(R/\mathrm{\Delta }r)^2(1\mathrm{cos}\alpha )+\mathrm{cos}\alpha ]=4Y[2i(R/\mathrm{\Delta }r)\mathrm{sin}(\alpha /2)\mathrm{cos}(\alpha /2)].$$
(103)
The stability condition $`Re[CD(AA^{})]0`$ is satisfied provided $`Ykl0`$, which rules out the artificially unstable solutions of the frozen PDE discussed above.
As a result, local stability analysis places no constraints on the algorithm. This may seem surprising because not even the analogue of a CFL condition on the time step arises but it can be understood in the following vein. The local structure of the code is implicit, since it involves 3 points at the upper time level. Implicit algorithms do not necessarily lead to a CFL condition. However, the algorithm is globally explicit in the way that evolution proceeds by an outward radial march from the origin. It is this feature that necessitates a CFL condition in order to make the numerical and physical domains of dependence consistent.
|
warning/0006/hep-ph0006340.html
|
ar5iv
|
text
|
# References
UCRHEP-T281
June 2000
Light Sterile Neutrinos from Large Extra Dimensions
Ernest Ma<sup>1</sup>, G. Rajasekaran<sup>1,2</sup>, and Utpal Sarkar<sup>1,3</sup>
<sup>1</sup> Department of Physics, University of California, Riverside, California 92521, USA
<sup>2</sup> Institute of Mathematical Sciences, Madras 600 113, India
<sup>3</sup> Physical Research Laboratory, Ahmedabad 380 009, India
## Abstract
An experimentally verifiable Higgs-triplet model of neutrino masses from large extra dimensions was recently proposed. We extend it to accomodate a light sterile neutrino which also mixes with the three active neutrinos. A previously proposed phenomenological model of four neutrinos (the only viable such model now left, in view of the latest atmospheric and solar neutrino-oscillation data) is specifically realized.
In the standard model of particle interactions, neutrinos $`\nu _i`$ belong in left-handed doublets $`(\nu _i,l_i)`$ under the electroweak $`SU(2)_L\times U(1)_Y`$ gauge group. Without right-handed singlet partners, they are massless but could obtain small Majorana masses through the effective dimension-5 operator
$$\frac{f_{ij}}{\mathrm{\Lambda }}(\nu _i\varphi ^0l_i\varphi ^+)(\nu _j\varphi ^0l_j\varphi ^+),$$
(1)
as the Higgs doublet $`(\varphi ^+,\varphi ^0)`$ acquires a nonzero vacuum expectation value ($`vev`$). The parameter $`\mathrm{\Lambda }`$ is an effective large mass scale.
Different models of neutrino mass are merely specific realizations of this operator in different extensions of the standard model. The usual approach is to identify the lepton-number violation with a very large scale, i.e. $`\varphi ^0=v<<\mathrm{\Lambda }`$. However, it is also possible that $`1/\mathrm{\Lambda }`$ is actually of the form $`m/M^2`$, so that the scale of lepton-number violation may be associated with $`m`$ which happens to be very small. Hence $`M`$ does not have to be very large, in which case the origin of neutrino mass may be tested directly in future high-energy colliders. In a recently proposed model in the context of large extra dimensions , exactly this very desirable feature is accomplished by adding a Higgs triplet $`(\xi ^{++},\xi ^+,\xi ^0)`$ to the standard model, as well as a scalar singlet $`\chi `$ which also exists in the bulk .
We now extend this proposal to include a light sterile neutrino $`\nu _s`$. (This is to be distinguished from other proposals where singlet fermions exist in the bulk which pair up with $`\nu _i`$ to become massive Dirac particles. We note that strong bounds on the fundamental scale in such theories have also been obtained .) The totality of present experimental evidence for neutrino oscillations calls for 1 sterile neutrino in addition to the 3 active ones. There are already a number of phenomenological scenarios for fitting all such data, but the latest results from the Super-Kamiokande Collaboration rule out the $`\nu _\mu \nu _s`$ explanation of the atmospheric data at the 99% confidence level, and also rule out the $`\nu _e\nu _s`$ explanation of the solar data at the 95% confidence level. This means that $`\nu _s`$ must be used in explaining the LSND result because a third mass-squared difference is required, and as far as we know, there is only one such phenomenological model that has been previously proposed . The salient feature of this model is the fast decay of $`\nu _s`$ into an antineutrino $`\overline{\nu }_i`$ and a massless Goldstone boson (the Majoron) corresponding to the spontaneous breaking of lepton number. We show in the following how this may naturally occur in the context of large extra dimensions.
The scalar singlet $`\chi `$ carries lepton number $`L=2`$ and its $`vev`$ is the source of all lepton-number violation in our four-dimensional world (called a 3-brane). The smallness of $`\chi `$ is due to the distance of our brane in the extra space dimensions from the source of the lepton-number violation located in another brane . Let $`y`$ denote a point in the extra dimensions. Our brane is located at $`y=0`$, whereas another 3-brane (the one providing the lepton-number violation) is at $`y=y_{}`$. We assume that they are separated by $`|y_{}|=r`$, where $`r`$ is the radius of compactification of the extra space dimensions, which is only a few $`\mu `$m in magnitude. The fundamental scale $`M_{}`$ in this theory is then related to the usual Planck scale $`M_P=2.4\times 10^{18}`$ GeV by the relation
$$r^nM_{}^{n+2}M_P^2,$$
(2)
where $`n`$ is the number of extra space dimensions. The scalar singlet $`\chi `$ exists in the bulk and could thus communicate between the 2 branes. For the source of lepton-number violation, one possibility is that a scalar field $`\sigma `$, carrying lepton number $`L=2`$, acquires a large $`vev`$ in the other brane. Assuming all mass parameters are of order $`M_{}`$, the field $`\sigma `$ has an interaction with $`\chi `$ given by
$$=\alpha M_{}^2d^4x^{}\sigma (x^{})\chi (x^{},y=y_{}),$$
(3)
where $`\alpha `$ is a parameter of order unity.
Once $`\sigma `$ acquires a $`vev`$, lepton number will be broken in the other brane. This will then act as a point source in the extra space dimensions for our world, so that the profile of $`\chi `$ is given by the Yukawa potential in the transverse dimensions :
$$\chi (y=0)=\sigma (y=y_{})\mathrm{\Delta }_n(r),$$
(4)
where
$$\mathrm{\Delta }_n(r)=\frac{1}{(2\pi )^{\frac{n}{2}}M_{}^{n2}}\left(\frac{m_\chi }{r}\right)^{\frac{n2}{2}}K_{\frac{n2}{2}}\left(m_\chi r\right),$$
(5)
$`K`$ being the modified Bessel function. Assuming that $`\sigma =M_{}`$, we then have the $`shining`$ of $`\chi `$ everywhere in our world corresponding to
$$\chi \frac{\mathrm{\Gamma }(\frac{n2}{2})}{4\pi ^{\frac{n}{2}}}\frac{M_{}}{(M_{}r)^{n2}}\frac{\mathrm{\Gamma }(\frac{n2}{2})}{4\pi ^{\frac{n}{2}}}M_{}\left(\frac{M_{}}{M_P}\right)^{2\frac{4}{n}},$$
(6)
where $`n>2`$ and $`m_\chi r<<1`$ have also been assumed. This $`shined`$ value of $`\chi `$ now appears as a boundary condition for our brane. In other words, the localized fields in our world must interact with $`\chi `$ in such a way that $`\chi `$ is unaffected.
Consider now the addition of a singlet left-handed sterile neutrino $`\nu _s`$. We propose the following 2 simple scenarios: (A) $`\nu _s`$ has $`L=+1`$, and (B) $`\nu _s`$ has $`L=1`$. In (A), the interaction Lagrangian involving $`\nu _i`$ and $`\nu _s`$ is given by
$$=f_{ij}[\xi ^0\nu _i\nu _j+\xi ^+(\nu _il_j+l_i\nu _j)/\sqrt{2}+\xi ^{++}l_il_j]+f_i\nu _s(\nu _i\eta ^0l_i\eta ^+)+h.c.,$$
(7)
where $`(\xi ^{++},\xi ^+,\xi ^0)`$ is a scalar triplet and $`(\eta ^+,\eta ^0)`$ is a scalar doublet, both carrying $`L=2`$. We define $`\chi z`$ and express the bulk field as
$$\chi =\frac{1}{\sqrt{2}}(\rho +z\sqrt{2})e^{i\phi }.$$
(8)
We assume that its behavior is not altered by the parameters in different branes. All such effects are already included in the boundary condition of Eq. (6). The lepton-number conserving interactions of $`\chi `$ with the other scalar fields in our world are then contained in
$`L`$ $`=`$ $`{\displaystyle }d^4x[hz(y=0)e^{i\phi (x)}(\overline{\xi }^0(x)\varphi ^0(x)\varphi ^0(x)\sqrt{2}\xi ^{}(x)\varphi ^+(x)\varphi ^0(x)+\xi ^{}(x)\varphi ^+(x)\varphi ^+(x))`$ (9)
$`+\mu z(y=0)e^{i\phi (x)}(\overline{\eta }^0(x)\varphi ^0(x)+\eta ^{}(x)\varphi ^+(x))+h.c.],`$
where $`\mu `$ is a mass parameter which could be of order $`M_{}`$ or the electroweak symmetry breaking scale. The self-interaction terms for the bulk scalar are now given by
$$V(\chi )=\lambda z(y)^2\rho (x,y)^2+\frac{1}{\sqrt{2}}\lambda z(y)\rho (x,y)^3+\frac{1}{8}\lambda \rho (x,y)^4.$$
(10)
This formulation has the virtue of universality, i.e. $`\lambda `$ is unchanged, but $`z`$ can change depending on where our brane is from the distant brane. It is also invariant under $`U(1)_L`$ : $`\nu _ie^{i\theta }\nu _i`$, $`\nu _se^{i\theta }\nu _s`$, $`\xi e^{2i\theta }\xi `$, $`\eta e^{2i\theta }\eta `$, $`\rho \rho `$, and $`\phi \phi 2\theta `$. The form of the potential $`V(\chi )`$ is that of the usual spontaneously broken U(1) theory and is independent of parameters in our brane.
The $`shining`$ of the field $`\chi `$ in our world induces a very weak lepton-number violating trilinear coupling of the Higgs triplet $`\xi `$ with the standard Higgs doublet $`\mathrm{\Phi }`$ , as well as the mixing of the Higgs doublets $`\eta `$ and $`\mathrm{\Phi }`$. In addition, the $`shined`$ value of $`\chi `$ supplies a Majorana mass term to the sterile neutrino through the interaction
$$L=f_sd^4xz(y=0)e^{i\phi (x)}\nu _s(x)\nu _s(x)+h.c.$$
(11)
From the Lagrangian of Eq. (9), it can easily be shown that $`\xi ^0`$ and $`\eta ^0`$ will have small $`vev`$s (say $`u`$ and $`w`$ respectively) which are proportional to $`z`$. Although the $`shined`$ value of $`\chi `$, i.e. $`z`$, comes from lepton-number violation in a distant brane and may not be determined in terms of the parameters entering in our world, the $`vev`$s of the other fields will be obtained in the usual way by minimizing the appropriate Higgs potential.
Consider the following Higgs potential containing the fields $`\xi `$, $`\eta `$, and $`\mathrm{\Phi }`$ with interaction terms involving $`\chi `$ in our brane:
$`V`$ $`=`$ $`m^2\mathrm{\Phi }^{}\mathrm{\Phi }+m_\xi ^2\xi ^{}\xi +m_\eta ^2\eta ^{}\eta +{\displaystyle \frac{1}{2}}\lambda _1(\mathrm{\Phi }^{}\mathrm{\Phi })^2+{\displaystyle \frac{1}{2}}\lambda _2(\xi ^{}\xi )^2+{\displaystyle \frac{1}{2}}\lambda _3(\eta ^{}\eta )^2+\lambda _4(\mathrm{\Phi }^{}\mathrm{\Phi })(\xi ^{}\xi )`$ (12)
$`+\lambda _5(\mathrm{\Phi }^{}\mathrm{\Phi })(\eta ^{}\eta )+\lambda _6(\xi ^{}\xi )(\eta ^{}\eta )+(hze^{i\phi }\xi ^{}\mathrm{\Phi }\mathrm{\Phi }+\mu ze^{i\phi }\eta ^{}\mathrm{\Phi }+h.c.)`$
Let $`\mathrm{\Phi }=v`$, $`\eta =w`$ and $`\xi =u`$, then the conditions for the minimum of $`V`$ are given by
$`v(m^2+\lambda _1v^2+\lambda _4u^2+\lambda _5w^2+2hzu)+\mu zw`$ $`=`$ $`0,`$ (13)
$`u(m_\xi ^2+\lambda _4v^2+\lambda _2u^2+\lambda _6w^2)+hv^2z`$ $`=`$ $`0,`$ (14)
$`w(m_\eta ^2+\lambda _5v^2+\lambda _6u^2+\lambda _3w^2)+\mu vz`$ $`=`$ $`0.`$ (15)
These equations tell us that the $`vev`$s of the fields $`\xi `$ and $`\eta `$ are small, and are given by
$$u\frac{hv^2z}{m_\xi ^2+\lambda _4v^2},w\frac{\mu vz}{m_\eta ^2+\lambda _5v^2},$$
(16)
where $`v\sqrt{m^2/\lambda _1}174`$ GeV as usual. \[In the above, we have neglected the term $`\xi ^{}\mathrm{\Phi }\eta `$ for simplicity. Its presence would only change the expressions for $`u`$ and $`w`$, not their proportionality to $`z`$. It also does not affect the composition of the Majoron in Eq. (21) to be given below.\] The $`4\times 4`$ neutrino mass matrix including the sterile neutrino in the basis $`(\nu _i,\nu _s)`$ is then of the form
$$_\nu =\left(\begin{array}{cc}2f_{ij}u& f_iw\\ f_iw& 2f_sz\end{array}\right),$$
(17)
where all the mass terms are naturally small, say of order 1 eV or less.
The kinetic-energy term of $`\chi `$ is
$$|_\mu \frac{1}{\sqrt{2}}(\rho +z\sqrt{2})e^{i\phi }|^2=\frac{1}{2}(_\mu \rho )^2+z^2(_\mu \phi )^2+\mathrm{},$$
(18)
which implies that $`\sqrt{2}z\phi `$ is the properly normalized massless Goldstone boson (the Majoron) from the spontaneous breaking of lepton number in the bulk. In the presence of the interaction terms of Eq. (9), other fields also participate in the spontaneous breaking of lepton number, hence the Majoron in our world becomes a combination of all these fields. The $`4\times 4`$ mass matrix in the basis $`(\mathrm{Im}\varphi ^0,\mathrm{Im}\xi ^0,\mathrm{Im}\eta ^0,z\phi )`$ is given by
$$\left(\begin{array}{cccc}4hzu\mu zw/v& 2hzv& \mu z& 2huv\mu w\\ 2hzv& hzv^2/u& 0& hv^2\\ \mu z& 0& \mu zv/w& \mu v\\ 2huv\mu w& hv^2& \mu v& huv^2/z\mu wv/z\end{array}\right).$$
(19)
Diagonalizing this matrix, we get one massless eigenstate
$$\frac{v\mathrm{Im}\varphi ^0+2u\mathrm{Im}\xi ^0+w\mathrm{Im}\eta ^0}{\sqrt{v^2+4u^2+w^2}},$$
(20)
which becomes the longitudinal component of the $`Z`$ boson, and another one which is the physical Majoron field:
$$N^{\frac{1}{2}}\left[v(w^2+2u^2)\mathrm{Im}\varphi ^0+u(v^2w^2)\mathrm{Im}\xi ^0+w(v^2+2u^2)\mathrm{Im}\eta ^0+z(v^2+w^2+4u^2)z\phi \right],$$
(21)
where $`N`$ is a normalization constant. The Majoron coupling matrix now becomes
$$\frac{1}{\sqrt{N}}\left(\begin{array}{cc}2f_{ij}u(v^2w^2)& f_iw(v^2+2u^2)\\ f_iw(v^2+2u^2)& 2f_sz(v^2+w^2+4u^2)\end{array}\right)$$
(22)
In the limit $`v\mathrm{}`$, the Majoron coupling matrix and the neutrino mass matrix are simultaneously diagonalized, in which case the Majoron will have only diagonal couplings. For finite $`v`$, the off-diagonal Majoron couplings to the neutrino mass eigenstates are suppressed by the factor $`(w^2+2u^2)/v^2`$, hence neutrino decay rates are very small and phenomenologically unimportant.
In scenario (B), $`\nu _s`$ has $`L=1`$. Hence the scalar doublet $`(\eta ^+,\eta ^0)`$ in Eq. (7) now carries no lepton number, i.e. $`L=0`$. To distinguish it from the usual Higgs doublet $`(\varphi ^+,\varphi ^0)`$, we add a discrete $`Z_2`$ symmetry, such that $`\nu _s`$ and $`\eta `$ are odd and all other fields are even. Instead of the $`e^{i\phi }\nu _s\nu _s`$ term in Eq. (11), we now have $`e^{i\phi }\nu _s\nu _s`$. Instead of the $`\mu ze^{i\phi }\eta ^{}\mathrm{\Phi }`$ term in Eq. (12), we now have $`\mu ^2\eta ^{}\mathrm{\Phi }`$ which breaks the $`Z_2`$ discrete symmetry softly, and as such, the parameter $`\mu `$ can be small. \[We have neglected the term $`h^{}ze^{i\phi }\xi ^{}\eta \eta `$ for simplicity. It does not affect the composition of the Majoron in Eq. (25) to be given below.\]
The equation of constraint for $`u`$ remains the same as Eq.(14), whereas the $`\mu zw`$ term in Eq. (13) is replaced by $`\mu ^2w`$, and the $`\mu vz`$ term in Eq. (15) is replaced by $`\mu ^2v`$. Hence $`u`$ is the same as in Eq. (16), and
$$w\frac{\mu ^2v}{m_\eta ^2+\lambda _5v^2}.$$
(23)
To obtain $`w1`$ eV, we need $`\mu 1`$ MeV for $`m_\eta 1`$ TeV.
The $`4\times 4`$ mass matrix in the basis $`(\mathrm{Im}\mathrm{\Phi }^0,\mathrm{Im}\xi ^0,\mathrm{Im}\eta ^0,z\phi )`$ now becomes
$$\left(\begin{array}{cccc}4hzu\mu ^2w/v& 2hzv& \mu ^2& 2huv\mu w\\ 2hzv& hzv^2/u& 0& hv^2\\ \mu ^2& 0& \mu ^2v/w& 0\\ 2huv\mu w& hv^2& 0& huv^2/z\end{array}\right),$$
(24)
and the physical Majoron is
$`N^{\frac{1}{2}}\left[2u^2v\mathrm{Im}\mathrm{\Phi }^0+u(v^2+w^2)\mathrm{Im}\xi ^02u^2w\mathrm{Im}\eta ^0+z(v^2+w^2+4u^2)z\phi \right]`$
$`(u\mathrm{Im}\xi ^0+z^2\phi )/\sqrt{u^2+z^2}.`$ (25)
Because $`\nu _s`$ has $`L=1`$ instead of $`L=+1`$, the Majoron coupling matrix is no longer proportional to the neutrino mass matrix even in the limit of $`v\mathrm{}`$. The latter remains the same as given by Eq. (17), but the former now differs from Eq. (22), i.e.
$$\frac{1}{\sqrt{u^2+z^2}}\left(\begin{array}{cc}2f_{ij}u& 0\\ 0& 2f_sz\end{array}\right).$$
(26)
When expressed in the basis of neutrino mass eigenstates, there are now unsuppressed off-diagonal terms. Hence neutrino decay is fast and the phenomenological model of Ref. is realized.
We advocate thus the $`4\times 4`$ neutrino mass matrix of Eq. (17), where the diagonal $`\nu _s`$ entry is of order a few eV. The $`3\times 3`$ submatrix spanning the $`\nu _e`$, $`\nu _\mu `$, and $`\nu _\tau `$ neutrinos is used to accommodate the atmospheric and solar neutrino data, and the mixing between $`\nu _s`$ and $`\nu _i`$ is used to fit the LSND data . Fast decay of $`\nu _s`$ into $`\overline{\nu }_i`$ and the Majoron allows us to evade the constraints of the CDHSW accelerator experiment , as explained in Ref. . This model can be tested in present and future solar-neutrino experiments by the observation of antineutrinos (instead of neutrinos) from the sun. Details are already given in Ref. .
In conclusion, we have shown how a light sterile neutrino may be naturally accommodated in the context of a theory of large extra space dimensions where a scalar field $`\chi `$ in the bulk carries lepton number which is spontaneously broken in a distant brane. The effects on our world are a small $`vev`$ for $`\chi `$ and its associated massless Goldstone boson (the Majoron). This allows us to have a naturally light $`4\times 4`$ mass matrix including the 1 sterile and the 3 active neutrinos. Two simple scenarios are discussed, one of which allows the fast decay of a heavier neutrino into a lighter antineutrino and the Majoron, as previously proposed . In view of the latest experimental results which exclude $`\nu _\mu \nu _s`$ in atmospheric and $`\nu _e\nu _s`$ in solar neutrino-oscillation data, this is the only viable such model now left.
Acknowledgement. This work was supported in part by the U. S. Department of Energy under Grant No. DE-FG03-94ER40837. G.R. and U.S. also thank the UCR Physics Department for hospitality.
|
warning/0006/quant-ph0006013.html
|
ar5iv
|
text
|
# Information, disturbance and Hamiltonian quantum feedback control LA-UR-00-1937
## I Introduction
With experimental advances, particularly in the fields of cavity QED and ion trapping , it is possible to observe individual quantum systems in real time, and it is therefore natural to consider the possibility of controlling such systems in real time using feedback . As a special case, real-time Markovian quantum feedback has been analyzed and implemented experimentally in certain quantum optical systems . Feedback control is invaluable in macroscopic applications, and as a consequence there is a vast body of literature devoted to classical feedback control. While results from classical control theory may be applied fruitfully to the quantum domain in certain special cases , these are not adequate in general, primarily because quantum measurement is quite different in nature from classical measurement, in that it has the capacity to disturb the system under observation . As a result, the development of optimal quantum control strategies requires optimizing over possible measurement strategies, which is unnecessary in classical control.
In feedback control the dynamics of a system is manipulated by using information obtained about the system through measurement. The goal is usually to maintain a desired state or dynamics in the presence of noise. A central problem of feedback control theory is the development of algorithms to achieve this goal. The approach to controller design that we consider here is to examine the measurement and feedback steps separately, thereby splitting the feedback control problem into two parts. One can then consider optimizing desirable properties of these parts separately under suitable constraints. If one allows the strength of either measurement or Hamiltonian feedback to be infinite, then any control objective can be achieved perfectly (this will be shown below once we have made these concepts of strength more precise). A constraint on strength is therefore the minimal constraint under which the problem of quantum feedback control is non-trivial, and this is the constraint we employ here.
The action of optimizing for the feedback and measurement independently ignores the possibility that truly optimal solutions may require considering both together. We will also simplify the problem by considering the optimization at each time step separately. This assumes that it is never desirable to perform worse at the current time in order to perform better at some future time. The approach we take here is therefore not aimed at finding a globally optimal solution given a set of constraints. However, the expectation is that the concepts we introduce here provide a simple systematic approach which one can expect to produce good results, and provide an insight into the kind of measurement processes which are desirable in feedback control.
For the feedback step, we consider the question of the effectiveness of the control by defining a cost function. Since one is interested in controlling the dynamics of a given quantum system (usually in the presence of some unavoidable source of environmental noise), one can specify the objective by specifying the most desired state for the system at each instant. The โcostโ function is then the sum of the distances of the state of the system from the desired state at each point in time, for some suitable measure of distance. Finding the optimal control strategy then consists in minimizing the cost function, under suitable constraints for the strength of the feedback. Note that this is somewhat different from the standard approach taken in modern classical control theory , and more similar to the approach taken in the new techniques of โpostmodernโ classical control . In modern classical control (e.g. LQG theory) one usually optimizes a โtotalโ cost function obtained from a suitably weighted sum of the cost function defined here, and another cost function intended to capture the cost of feedback strength.
We will restrict ourselves to control objectives such that the desired state at each time (the target state) is pure. Impurity (mixing) merely signifies that our knowledge of the quantum system is less than maximal, which is by assumption undesirable.
In considering the optimality of the measurement step, rather than attempting to find a measurement which explicitly optimizes the cost function, we define concepts of information and disturbance, motivated by the feedback control problem. We then consider finding measurements which maximize the information and minimize the disturbance. We find that in general these two targets are mutually exclusive, in striking contrast to classical control theory. This implies the existence of a trade-off between information and disturbance in quantum feedback control.
Since we focus on continuous feedback control, and many readers will be familiar with generalized measurements but unfamiliar with the formalism of continuous quantum measurement, we describe in the next section how continuous observation is formulated within the language of generalized measurements. In Section III we define the concept of the strength of a measurement, required as a minimal constraint for the feedback control problem. In Section IV we discuss in detail the division of feedback control into โpureโ measurement and Hamiltonian feedback, and consider what may be achieved when there is no limitation on the strength of either. We also discuss what may be achieved in this case both without feedback and with measurement-only feedback. In Section V, we consider the measurement process, define concepts of information and disturbance, and consider minimizing the disturbance and maximizing the information. In Section VI we examine the Hamiltonian feedback and obtain Hamiltonians which minimize the instantaneous cost function. In Section VII we implement the feedback control of a two-state system, showing how the ideas presented in the previous sections manifest in the performance of the control algorithm. Section VIII summarizes and concludes.
## II Continuous observation and generalized measurements
We will concern ourselves primarily with continuous-time quantum feedback control, in which a system is observed continuously, and the results of the measurements (the measurement record) used to continuously alter the Hamiltonian of the system to effect control. We now discuss how continuous observation may be described within the language of generalized quantum measurements, implemented as positive operator valued measures (POVMโs).
Continuous measurements on a quantum system generate a measurement record that is a continuous-time stochastic process, which may be either a (Gaussian) Wiener process or a Poisson process . For a given physical system, these two kinds of processes will result from making different measurements, for example photon counting and homodyne detection of optical beams.
The key ingredient in describing continuous measurements is that during an infinitesimal time step $`dt`$, the information obtained by the observer must scale as $`dt`$, so that one can take the continuum limit and obtain a sensible answer. This may be realized by defining a POVM, given by $`\mathrm{\Omega }_\alpha ^{}\mathrm{\Omega }_\alpha ๐\alpha =1`$, to describe the result of an observation in the time interval $`dt`$ by
$$\mathrm{\Omega }_\alpha =\left(\frac{\pi }{2dt}\right)^{1/4}e^{kdt(Q\alpha )^2},$$
(1)
where $`Q`$ is a arbitrary operator for the system under observation, $`\alpha `$ takes all values on the real line, and $`k`$ is a positive real constant. For reasons that will be made clear in the next section, we will only need to be concerned with the case in which $`Q`$ is Hermitian, so that $`Q`$ may be referred to as an observable, and we will assume this in what follows. Note that each $`\mathrm{\Omega }_\alpha `$ is a weighted sum of projectors onto the eigenbasis of $`Q`$, where the weighting is peaked at $`\alpha `$. Thus each application of the one of the $`\mathrm{\Omega }^{}s`$ provides some information about the observable $`Q`$. However, as $`dt`$ tends to zero, this information also tends to zero, since the $`\mathrm{\Omega }^{}s`$ become increasingly broad over the eigenstates of $`Q`$. Calculating the measurement result in the interval $`dt`$ at time $`t`$, and denoting this as $`dy(t)`$, we have
$$dy(t)=4kQdt+\sqrt{2k}dW,$$
(2)
where $`dW`$ is the Wiener increment for the interval $`dt`$. Using this, one can obtain the stochastic evolution of the quantum state under this measurement process, referred to as a quantum trajectory, and this is given by the Stochastic Master Equation (SME)
$`d\rho `$ $`=`$ $`i[H,\rho ]dtk[Q,[Q,\rho ]]dt`$ (4)
$`+(Q\rho +\rho Q2\text{Tr}[Q\rho ]\rho )\sqrt{2k}dW.`$
where $`H`$ gives the system evolution in the absence of the measurement. We can also readily obtain the non-selective evolution, in which the measurement results are ignored, and this is given by
$`\rho (t+dt)`$ $`=`$ $`i[H,\rho ]dt+{\displaystyle \mathrm{\Omega }_\alpha \rho \mathrm{\Omega }_\alpha ^{}๐\alpha }`$ (5)
$`=`$ $`i[H,\rho ]dtk[Q,[Q,\rho (t)]]dt.`$ (6)
When $`H`$ commutes with $`Q`$ this evolution leads to a diagonalization of $`\rho `$ in the basis of $`Q`$, as one would expect for a measurement of $`Q`$. Similarly, integrating the SME in this case, one finds that the result in the long time limit is a projection onto one of the eigenstates of $`Q`$. Such a POVM realizes a continuous measurement of the operator $`Q`$, such that the measurement record is a Wiener process.
One can also define a POVM to provide continuous observation in which the measurement record is a Poisson process. Since this requires only one of the two possible outcomes at each interval $`dt`$, the POVM consists of only two measurement operators:
$`\mathrm{\Omega }_0`$ $`=`$ $`1\frac{1}{2}kQ^2dt,`$ (7)
$`\mathrm{\Omega }_1`$ $`=`$ $`Q\sqrt{kdt}.`$ (8)
That this gives a Poisson process can be seen by considering the probabilities for the outcomes 0 and 1, which are $`1kQdt`$ and $`kQdt`$, respectively. Result 1 therefore corresponds to an Poisson โeventโ, which happens occasionally, and 0 to the absence of one. The SME corresponding to the measurement process is different from that corresponding to the Wiener measurement, but the non-selective evolution is identical. Physically, the non-selective evolution is fixed by choosing the interaction of the system with the environment which is mediating the measurement, and the trajectory, whether Poisson or Wiener, is selected by how one chooses to measure the environment so as to extract the information about the system. In fact, by taking a suitable unitary transformation of the Poisson measurement operators, and taking the appropriate limit in which there are many events in each interval $`dt`$, one can obtain the Wiener process measurement from the Poisson measurement, and so the first can be regarded as a special case of the second. This is discussed in detail in .
The point we wish to note here is that regardless of how one chooses the trajectory, a continuous measurement of an observable $`Q`$ is given by a POVM in which all the measurement operators $`\mathrm{\Omega }_\alpha `$ are positive operators, diagonal in the basis of $`Q`$, and one must merely be careful to choose the form of these operators with respect to $`dt`$ so as to provide a sensible continuum limit.
## III The Strength of a Measurement
Clearly the more accurate the measurements of the observer, the more information she is able to obtain, and better able she is to choose feedback to effectively control the system. However, in general, more accurate measurements require more resources. In treating quantum feedback control, it is sensible to consider a restriction on available resources, and hence a restriction on measurement accuracy. To treat this quantitatively, one must introduce a sufficiently precise notion of the accuracy, or strength, of a quantum measurement.
For the purposes of feedback control, since it is the final state resulting from measurement that the observer must act upon with feedback, it is the observers information about this final state which is relevant. Intuitively, one can therefore think of stronger measurements as providing, on average, final states which are more pure (or, alternatively, have a smaller von Neumann entropy) than weaker measurements. When considering continuous observation, in the absence of any noise sources, an initially impure state is continually purified. In this case the strength of the measurement can be thought of as being proportional to the rate of this purification. Note that this concept of information extraction by a measurement is quite different from that usually considered in quantum information theory. There, authors have been concerned about the information that a measurement provides about the initial state of the system (the state immediately before the measurement) , whereas in our case it is the information about the final state which is important.
We will not need an explicit definition for measurement strength here, since we will only require two properties of measurement strength which we will motivate below. However, we will give an example of an explicit definition which satisfies these two properties. To motivate the first property, we note that as we have defined it so far, it is clear that the strength of a measurement in some sense characterizes the average rank of the operators $`\mathrm{\Omega }_m`$ which make up the associated POVM ($`_m\mathrm{\Omega }_m^{}\mathrm{\Omega }_m=1`$). If all the $`\mathrm{\Omega }_m`$ are rank one, then one always obtains a pure final state, and therefore complete information, regardless of the initial state. The higher the rank of the projectors, the higher in general will be the von Neumann entropy for a fixed initial state. The first property we will require is that measurements that consist of rank one projectors should have maximum strength (for measurements on a system of a given dimension).
For the remainder of this paper, we will refer to measurements for which at least one of the $`\mathrm{\Omega }_m`$ are rank one as infinite strength measurements. This terminology is natural in the context of continuous observation, since in order to provide rank one projections in a finite time from a continuous measurement, one would have to take the limit $`k\mathrm{}`$ in Eq.(4). However, we wish to stress that our use of this terminology is not intended to imply that any explicit definition of measurement strength should necessarily take this value for these kinds of measurements.
The second property we wish to impose is that strength be invariant under unitary transformations of the measurement operators. To motivate this property, one can consider a device which measures the spin of a two state system. One would expect such a device to provide the same strength of measurement regardless of how it is oriented in space. Since spatial rotation covers all unitary transformations for a spin-half, for this system strength should be invariant under all unitary transformations of the $`\mathrm{\Omega }_n`$. We will explicitly consider the spin-half system later.
To provide an example of an explicit definition of measurement strength for single-shot measurements on finite dimensional systems, one can first consider the average uncertainty after the measurement result is known. Using the von Neumann entropy, for a measurement described by $`_n\mathrm{\Omega }_n^{}\mathrm{\Omega }_n=1`$, this is
$$u_\text{V}(\rho )=\underset{n}{}\text{Tr}[\mathrm{\Omega }_n\rho \mathrm{\Omega }_n^{}\text{ln}(\mathrm{\Omega }_n\rho \mathrm{\Omega }_n^{}/\text{Tr}[\mathrm{\Omega }_n\rho \mathrm{\Omega }_n^{}])],$$
(9)
where $`\rho `$ is the initial state of the system. Using the purity as an alternative measure of uncertainty we have
$$u_\text{p}(\rho )=1\underset{n}{}\frac{\text{Tr}[(\mathrm{\Omega }_n\rho \mathrm{\Omega }_n^{})^2]}{\text{Tr}[\mathrm{\Omega }_n\rho \mathrm{\Omega }_n^{}]}.$$
(10)
To obtain definitions of measurement strength that satisfy our two properties we can use the following functions of these uncertainties:
$`s_\text{v}`$ $`=`$ $`{\displaystyle \frac{1}{u_\text{V}(I/N)}}{\displaystyle \frac{1}{\mathrm{log}(N)}},`$ (11)
$`s_\text{p}`$ $`=`$ $`{\displaystyle \frac{1}{u_\text{p}(I/N)}}{\displaystyle \frac{N}{(N1)}},`$ (12)
in which $`N`$ is the dimension of the system being measured. These particular definitions have the additional property that they tend to infinity for von Neumann measurements, but no others. It is also a fairly simple matter to write explicit definitions of measurement strength that tend to infinity for measurements that contain at least one rank one projector.
Definitions of measurement strength for single-shot measurements may be extended to continuous measurements by using the rate of uncertainty reduction. Using the explicit definitions given above (Eq.(11) and Eq.(12)), it is straightforward to calculate this rate from Eq.(4) and the Ito rules for stochastic differential equations :
$`{\displaystyle \frac{d}{dt}}s_\text{v}`$ $`=`$ $`{\displaystyle \frac{4k}{N\mathrm{log}(N)^2}}\left(\text{Tr}[Q^2]+3\text{Tr}[Q]^2\right)`$ (13)
$`{\displaystyle \frac{d}{dt}}s_\text{p}`$ $`=`$ $`8k{\displaystyle \frac{N^2}{(N1)^2}}\text{Tr}[Q^2].`$ (14)
In considering feedback control, measurement strength is a particularly important concept because it is a constrained resource; stronger measurements are in general more expensive. A particular example is the measurement of position by the reflection of a laser-beam , a technique used in the atomic force microscope. In that case it is the laser power on which the measurement strength depends. Hence it is reasonable to consider optimization of the measurement under the assumption that measurement strength is fixed.
## IV Measurement and Feedback
In classical feedback control, it is natural to consider the measurement process as being qualitatively different from the feedback process. In particular they may be distinguished by the fact that the measurement in each time step involves no change to the system Hamiltonian, and the feedback step provides no information. In quantum feedback, since measurement has the ability to affect the dynamics in ways that in classical mechanics would have to be attributed to a Hamiltonian, the distinction is not as fundamental. However, in the vast majority of quantum feedback schemes considered to date, it is some set of parameters describing the system Hamiltonian which are under the observerโs control. This is motivated by practical considerations, since it is as of yet easiest experimentally to externally control aspects of the Hamiltonian. In this case the feedback step involves no measurement, and the observation and feedback processes may be regarded as qualitatively different, as in the classical theory. In view of this, the polar decomposition theorem motivates some definitions.
By Krausโs representation theorem , every valid quantum evolution (a quantum operation) may be written as a POVM given by a set of operators $`\mathrm{\Omega }_n`$, where the probability of each outcome is $`P(n)=\text{Tr}[\mathrm{\Omega }_n^{}\mathrm{\Omega }_n\rho ]`$ and the state resulting from each outcome is $`\rho _n=\mathrm{\Omega }_n\rho \mathrm{\Omega }_n^{}/P(n)`$. The only constraint on the $`\mathrm{\Omega }_n`$โs is that $`_n\mathrm{\Omega }_n^{}\mathrm{\Omega }_n=1`$. However, from the polar decomposition theorem, each of the operators $`\mathrm{\Omega }`$ may be written as the product of a unitary operator and a positive operator, so that
$$\mathrm{\Omega }_n=U_n\sqrt{\mathrm{\Omega }_n^{}\mathrm{\Omega }_n}.$$
(15)
This provides a natural decomposition of a general quantum operation in terms of measurement and feedback. Consider first the action of the unitary operators. By themselves they do not describe the acquisition of information, and in that sense they do not describe a measurement. This can be seen from the fact that a unitary operator does not change the von Neumann entropy of any state it acts upon, and consequently extracts no information. However, unitary operations are precisely the kind that can be applied by Hamiltonian feedback. Hence, the unitary operators appearing in the polar decomposition may be thought of as characterizing purely the feedback part of the quantum operation. Note that we have written the polar decomposition so that the action of the unitary operator follows after the action of the positive operator, being a necessary condition for feedback.
Conversely, the positive operators characterize the acquisition of information. They may always be written as a weighted sum of projectors, and therefore thought of as providing partial information about the states in the basis in which they are diagonal. When they correspond to rank 1 projectors, they provide complete information, in that the final state is pure. Since the unitary part has been factored out to obtain the positive operators, we may regard these operators as representing pure measurement; the change induced in the quantum state is only that which is strictly necessary in order provide the information obtained during the measurement. We note that this decomposition of measurements into unitary and positive operators has been considered before in the context of measurements of the first and second kind .
From this it is clear that every quantum evolution can be realized by a measurement in which the measurement operators are positive, followed by a feedback step in which the Hamiltonian is chosen to depend upon the measurement result. We see that the observation of a single observable, considered in section II, corresponds to the special case in which all the positive operators forming the POVM are mutually commuting.
Under the above definitions damping processes, such as cavity decay and Brownian motion, are not considered pure measurements; they are viewed as equivalent to a fixed combination of measurement and feedback. Since the object of feedback control is to limit the deviations of a system from a desired state (or more, generally, from a particular evolution - which means merely that the target state changes with time), feedback control is essentially a damping process (toward the target state).
The polar decomposition theorem therefore fits snugly with the structure of Hamiltonian feedback, but it is nevertheless important to realize that this is not the only feedback process that may be considered in quantum mechanics. First note that the product of two positive operators need not be positive. Hence the evolution resulting from a sequence of pure measurements as defined above will in general be equivalent to a single pure measurement followed by some Hamiltonian evolution (ie. both measurement and Hamiltonian feedback). This is an illustration of the fact that quantum measurements involve โactiveโ transformations of the states, as opposed to the โpassiveโ measurements of classical physics .
Consider now the full evolution of a system under Hamiltonian feedback control in a single infinitesimal time step $`dt`$, with initial state $`\rho `$. Since all dynamical processes commute to first order, one can treat even continuous feedback control as alternating steps consisting of measurement and feedback. This is consistent with the general approach of this paper which is to consider the two steps separately. The system evolves under its own โfreeโ Hamiltonian, $`H_0`$, (which in many cases will be the desired evolution), and is affected by a source of environmental noise, which can be described by the non-selective evolution generated by a POVM. The measurement is also performed, and the feedback evolution applied. For a given measurement result $`n`$, we may write the full evolution as
$$\stackrel{~}{\rho }_n=e^{i(H_n+H_0)dt}P_n(\underset{j}{}W_j\rho W_j^{})P_ne^{i(H_n+H_0)dt}$$
(16)
where the tilde indicates that we have not bothered to normalize the final state, and $`W_j`$ are the operators describing the (undesirable) action of the environment. Since all the operators always commute to first order in $`dt`$, we have combined the free Hamiltonian with the feedback Hamiltonian in the exponential. The task of feedback control is to choose operators $`P_n`$ and $`H_n`$ such that the evolution is closest to the desired evolution. Before we consider this for Hamiltonian feedback, let us examine what can be done in the absence of the conditional unitaries, using measurement alone, and the difference between the two kinds of feedback.
By the definition above, using measurement alone one is restricted to POVMโs in which all the measurement operators are positive, along with some overall unitary evolution independent of the measurement results. Now, to evaluate the efficacy of the control procedure, we must have a โcost functionโ which measures how well we have achieved the control objective, as discussed above in the introduction. Since we have a desired โtargetโ state, $`\sigma =|\psi _\text{T}\psi _\text{T}|`$, in mind at some final time (to be achieved following a single measurement, or a series of measurements), sensible cost functions will provide a measure of how close the final state, $`\rho _\text{f}`$, is to the target state. A number of measures are possible, such as the inner product ($`\text{Tr}[\rho _\text{f}\sigma ]`$), the fidelity ($`\text{Tr}[\sqrt{\sigma ^{1/2}\rho _\text{f}\sigma ^{1/2}}]`$), or the destinguishability ($`(1/2)\text{Tr}[\rho _\text{f}\sigma ]`$). Since we are interested only in target states which are pure, the Fidelity is simply the square root of the inner product, so that they provide equivalent optimization problems. Throughout this paper we will use these as the quantities to be optimized.
Now, the final state resulting from averaging the results of a single pure measurement is given by
$$\rho _\text{f}=\underset{n}{}P_n\rho P_n.$$
(17)
Since $`P_n=P_n^{}`$, Andoโs result states that $`\rho _\text{f}`$ is always majorized by $`\rho `$, which means that the eigenvalues of $`\rho _\text{f}`$ are at least as evenly distributed as the eigenvalues of $`\rho `$. This means that the von Neumann entropy of $`\rho _\text{f}`$ is always at least as large as the entropy of $`\rho `$. Another way of putting this is that each eigenvalue of $`\rho _\text{f}`$ is some weighted average of one or more of the eigenvalues of $`\rho `$.
It follows almost immediately from the above results, that the fidelity of the final state cannot be any larger than the maximum eigenvalue, $`\lambda _{\text{max}(\rho )}`$, of the initial state $`\rho `$. To see this we first note that since all the eigenvalues of the final state, $`\lambda _j`$ are a weighted average of the eigenvalues of $`\rho `$, none can be larger than the largest eigenvalue of $`\rho `$. Now, writing the fidelity in terms of the eigenvectors of $`\rho _\text{f}`$, $`|\varphi _j`$ we have
$$\psi _\text{T}|\rho _\text{f}|\psi _\text{T}=\underset{j}{}\lambda _j|\varphi _j|\psi _\text{T}|^2.$$
(18)
Since $`_j|\varphi _j|\psi _\text{T}|^2=1`$, the fidelity is merely a weighted average of the eigenvalues of $`\rho _\text{f}`$, which proves the result. In fact, choosing any basis $`|\psi _i`$, we obtain the probability distribution over these states as
$$\mu _i=\psi _i|\rho _\text{f}|\psi _i=\underset{j}{}\chi _{ij}\lambda _j,$$
(19)
where $`\chi _{ij}=|\varphi _j|\psi _i|^2`$. Since the matrix $`\chi _{ij}`$ satisfies $`_i\chi _{ij}=1`$ and $`_j\chi _{ij}=1`$, it is a doubly stochastic map, with the result that the vector $`\{\mu _i\}`$ is majorized by the vector $`\{\lambda _j\}`$, and hence the von Neumann entropy of the distribution over any set of basis states is always at least as large as the distribution over the eigenvectors. Another way of saying this is that diagonal elements of a matrix resulting from a unitary transformation performed on a diagonal matrix are always at least as uniformly distributed as the original elements (and almost always more so).
Clearly this result for the upper bound on the final fidelity also holds for repeated measurements, in which subsequent measurements are not conditioned on the results of previous measurements (i.e. for pure measurements with no feedback). However, it does not hold for sequences of conditional measurements. In this case the initial state seen by subsequent measurements cannot be written as the state given by averaging over the results of previous measurements, since each final state may have a different measurement performed on it.
It turns out that if we allow ourselves an infinite measurement strength, then the upper bound on the final entropy derived above can always be achieved in the limit of an infinite number of measurements. To see this one simply follows the procedure of Aharonov and Vardi, referred to as the โinverse quantum Zeno effectโ, developed in reference . Consider first an initial pure state, $`|\psi `$, and the projector, $`P_\epsilon =|\epsilon \epsilon |`$ onto the state
$$|\epsilon =(\mathrm{cos}(\epsilon )|\psi +\mathrm{sin}(\epsilon )|\psi _\text{T})$$
(20)
For $`\epsilon =0`$ this is the initial state, and for $`\epsilon =\pi /2`$ this is the final state. For any value in between this state represents a rotation through an angle $`\epsilon `$ from the initial state to the final state. If $`P_\epsilon =|\epsilon \epsilon |`$ makes up one of the operators describing the measurement, the probability of failing to obtain the state $`|\epsilon `$ is
$$P(\epsilon )=\mathrm{sin}^2(\epsilon )=\epsilon ^2+\mathrm{}$$
(21)
If we succeed in obtaining the state $`|\epsilon `$, then in that measurement step we have succeeded in rotating the state through $`\epsilon `$ toward the desired state. We can attempt to rotate the state through the full $`\pi /2`$ radians by choosing $`\epsilon =\pi /(2M)`$ and using $`M`$ measurements. Since the probability of failure at each step is then second order in $`(1/M)`$, while the number of steps scales only as $`M`$, as the number of steps tends to infinity, the total probability of failure tends to zero, and in this limit one achieves the desired rotation. To see that this achieves the upper bound when the initial state is mixed, we choose the projector so as to rotate the eigenvector of $`\rho `$ corresponding to the maximum eigenvalue to the desired state. For each measurement the corresponding POVM is then given, in general, by $`P_{\epsilon _i}+_l\mathrm{\Omega }_l^{}\mathrm{\Omega }_l=1`$, where $`P_{\epsilon _i}`$ is the projector for the $`i^th`$ measurement, and the $`\mathrm{\Omega }_l`$ are arbitrary. The final state may then be written
$`\rho _\text{f}={\displaystyle \underset{i}{\overset{M}{}}}_{\epsilon _i}\left[\lambda _1|11|\right]+{\displaystyle \underset{i}{\overset{M}{}}}_{\epsilon _i}\left[{\displaystyle \underset{n=2}{\overset{N}{}}}\lambda _n|nn|\right]`$ (22)
$`_{\epsilon _i}\left[A\right]=P_{\epsilon _i}AP_{\epsilon _i}+{\displaystyle \underset{l}{}}\mathrm{\Omega }_lA\mathrm{\Omega }_l^{}`$ (23)
where $`\rho |n=\lambda _n|n`$, with $`\lambda _1`$ the maximum eigenvalue, and $`A`$ is an arbitrary operator. As the number of measurements tends to infinity, the first term on the RHS of Eq.(22) becomes
$$\underset{M\mathrm{}}{lim}_{\epsilon _i}\lambda _1|11|=\lambda _1|\psi _\text{T}\psi _\text{T}|.$$
(24)
Since this term contributes $`\lambda _{\text{max}}(\rho )`$ to the fidelity, and since the other pure states making up the final density matrix cannot contribute negatively, the upper bound is achieved.
What happens when we allow ourselves infinite measurement strength, and sequences of measurements in which subsequent measurements are conditioned on previous results (i.e. measurement-only feedback)? In that case it is clear that the desired state can always be obtained with certainty; one begins by making a projection measurement in an arbitrary basis, which results in a set of pure states. Then the above procedure is used to rotate the resulting state to the target.
When infinite strength is available, measurement-only feedback is equivalent to Hamiltonian feedback, since both allow any state to be created. However, in many continuous feedback control applications, the strength of the measurement is unlikely to be so much stronger than either the environmental noise or the free system dynamics that it can be used in this fashion in place of Hamiltonian feedback . With measurements of finite strength the outcomes are necessarily random, so that Hamiltonian feedback cannot be simulated reliably. One can expect therefore that real applications will find the use of Hamiltonian feedback invaluable.
## V Measurement: Maximal Information and Minimal Disturbance
In Section III we introduced the concept of the information provided about the system by a measurement, and this involved specifically the information regarding the state resulting from the measurement, a definition motivated by feedback control. This in turn motivated the definition of the strength of a measurement (e.g. $`s_\text{v}`$ or $`s_\text{p}`$), important because it constitutes a natural constraint when considering the optimization of control strategies. However, the actual information provided by a given measurement is not only a function of the measurement strength, but also the state of the system immediately prior to the measurement. As a result, once the available measurement strength is known, one can ask how to optimize the information provided by the measurement given the current state of the system. This defines the concept of a measurement returning maximal information (for a fixed measurement strength).
In addition to providing information, quantum measurements can also introduce noise, a statement which we will now make precise. Consider first a classical system driven by noise. One can characterize the extent of the noise in some time interval by the increase in the entropy of the phase space probability distribution for the system state which is given by averaging over the noise realizations. This tells us how much we expect the noise to spread out the system in phase space in that time interval, and characterizes our uncertainty about the future state of the system resulting from the noise. Now consider a classical measurement. Since the initial state is uncertain (or else we would not need to make the measurement), the result of the measurement is random, and as a result oneโs state of knowledge changes in a random fashion. However, this random change should not be considered noise, since if one averages over all the possible measurement results (all the possible random changes) the probability distribution for the state of system remains unchanged. This is the sense in which classical measurement introduces no noise into the system.
Now consider a quantum system driven by noise. The equivalent of the phase space distribution is the density matrix. In the same manner one can characterize noise by the resulting increase in the von Neumann entropy of the density matrix resulting from averaging over the possible noise realizations. One can therefore characterize the noise introduced by a quantum measurement by calculating the increase in the von Neumann entropy (or alternatively the decrease in the purity) of the density matrix which results from averaging over the possible measurement results. While we saw above that in the classical case the measurement introduces no noise, this is not, in general true for quantum measurement. In terms of the von Neumann entropy, the excess noise introduced by a measurement is
$$N_\text{e}^\text{v}=S_\text{v}(\rho _\text{f})S_\text{v}(\rho ).$$
(25)
Defining it in terms of the purity we have
$$N_\text{e}^\text{p}=\text{Tr}[\rho ^2]\text{Tr}[\rho _\text{f}^2].$$
(26)
This makes precise the intended meaning of our initial statement that quantum measurements can introduce noise. Note that this has nothing to do with the Heisenberg uncertainty principle โ what concerns us here is the uncertainty of the future quantum state, and not the uncertainty of some set of observables for a given state. Recall that this is because the object of the control is the state of the system, and it is up to the observer to decide what the desired state is. Whether it be a minimal uncertainty state in the sense of the Heisenberg uncertainty principle is immaterial.
Let us consider first the question of minimizing the disturbance due to the measurement. Recall that for a pure measurement, the evolution given by averaging over the measurement results is given by Eq. (17), where all the $`P_n`$ are positive operators. Once again invoking Andoโs result, we have that the von Neumann entropy of the final state is never decreased by the measurement. Measurements which minimize noise are therefore the measurements which leave the von Neumann entropy unchanged. These measurements are in this sense most like classical measurements. A set of measurements satisfying this criterion are those in which all the $`P_n`$ commute with the initial density matrix. In this case we have
$$\rho _\text{f}=\underset{n}{}P_n\rho P_n=\underset{n}{}P_n^2\rho =\rho .$$
(27)
In the language of continuous measurements, since the operators $`P_n`$ are diagonal in the eigenbasis of the observable, this means choosing to measure an observable which shares an eigenbasis with the density matrix.
On a practical note, for continuous observation, measuring in the eigenbasis of the density matrix involves continuously changing the measured observable (note that such a process has been considered previously in the context of adaptive measurements ). In many situations this flexibility may be only partially available, or not at all. However, the above analysis indicates that for the purposes of noise minimization, one should choose the measured observable to be that in which the system is diagonal, or nearly diagonal, for the longest time during the period of control. In fact, this introduces the possibility that in certain cases it may be desirable to turn off measurement for periods in which the system occupies states which have large off-diagonal elements in the eigenbasis of the observable. Of course, the resulting noise reduction would have to be balanced against the accompanying loss of information.
Maximizing the information for a fixed measurement strength is a much more difficult problem. Here we will examine a specific example for the continuous measurement of a two-state system. In the formulation of continuous measurements that was discussed in Section II, we used measurement operators where each was a sum over an infinite number of projectors. For a two state system it is possible to obtain the same result (i.e. the same continuous measurement driven by Gaussian noise) by using a formulation with only two measurement operators where each is the sum over only two projectors. To obtain a continuous measurement of a given observable the POVM is given by $`\mathrm{\Omega }_0^2+\mathrm{\Omega }_1^2=1`$ where the measurement operators are
$`\mathrm{\Omega }_0`$ $`=`$ $`\sqrt{\kappa }|00|+\sqrt{1\kappa }|11|,`$ (28)
$`\mathrm{\Omega }_1`$ $`=`$ $`\sqrt{\kappa }|11|+\sqrt{1\kappa }|00|`$ (29)
in which $`\kappa =1/2+\sqrt{kdt}`$, and $`|0`$ and $`|1`$ are the eigenstates of the observable. In each time step $`dt`$ this produces one of two results. The sum of these, in a time interval $`\mathrm{\Delta }t=Ndt`$ in which $`N`$ results are obtained, is naturally governed by the binomial distribution. In the limit of large $`N`$ (and infinitesimal $`\mathrm{\Delta }t`$) this tends to a Gaussian, and one obtains the measurement record (Eq.(2)) and SME (Eq.(4)) given in Section II, where the measured observable $`Q=|00||11|`$. We can alternatively think of this measurement as a single-shot measurement, and in that case $`\kappa `$ can take any value between zero and one. Note that when $`\kappa =0`$ or $`\kappa =1`$ the measurement is one of infinite strength. As $`\kappa `$ becomes closer to $`1/2`$, the strength reduces, and for $`\kappa =1/2`$ the measurement provides no information.
We can obtain measurements of all possible observables by applying to the measurement operators an arbitrary rotation over the Bloch sphere, given by the unitary transformation
$`U(\theta ,\varphi )|0`$ $`=`$ $`\mathrm{cos}(\theta /2)|0+e^{i\varphi }\mathrm{sin}(\theta /2)|1`$ (30)
$`U(\theta ,\varphi )|1`$ $`=`$ $`\mathrm{cos}(\theta /2)|1e^{i\varphi }\mathrm{sin}(\theta /2)|0.`$ (31)
Recall that this unitary transformation of the measurement operators preserves the measurement strength as defined in section III. Without loss of generality, we can choose the initial density matrix to be diagonal, and write it as $`\rho =p|00|+(1p)|11|`$. One can then obtain an analytic expression for the final average purity, which is given by
$$I_\text{f}^\text{p}1u_\text{p}=\underset{n=0}{\overset{1}{}}\frac{\text{Tr}[(U\mathrm{\Omega }_n^{}U^{}\rho U\mathrm{\Omega }_nU^{})^2]}{\text{Tr}[U\mathrm{\Omega }_n^{}U^{}\rho U\mathrm{\Omega }_nU^{}]}.$$
(32)
This expression is fairly complex, and we will not need it here. (For a detailed analysis of this expression, including analytic expressions for general two-outcome measurements on two-state systems, the reader is referred to ). It is explicitly independent of $`\varphi `$, as one would expect, since it is $`\theta `$ alone which gives the angle (on the Bloch sphere) between the basis in which the density matrix is diagonal, and the basis of the measured observable. The final average purity is then explicitly dependent on the three parameters, $`p`$, $`k`$ and $`\theta `$, and we are concerned with maximizing this with respect to $`\theta `$. When the measurement is non-trivial ($`\kappa 0.5`$), the strength of the measurement is finite ($`\kappa 0,1`$) and the initial state is impure ($`p(0,1)`$), one finds that the location of the maximum is independent of $`p`$ and $`k`$, and occurs for $`\theta =\pi /2`$. This means that on average the maximum information is obtained about the final state (for fixed measurement strength) when the basis of the measured observable is maximally different from the eigenbasis of the density matrix (ie. if the density matrix is a mixture of $`\sigma _z`$ eigenstates, then one should measure $`\sigma _x`$ or $`\sigma _y`$).
We see then, that at least for a two-state system, the minimal disturbance is obtained when the measured observable has the same eigenbasis as the density matrix, and the maximal information is obtained when the two bases are maximally different. Thus we obtain the result that, at least for a two-state system, there is a trade-off between information and disturbance in quantum feedback control (in contrast to classical feedback control). This trade-off for finite strength quantum measurements is also of interest from a purely fundamental point of view, and this is explored in detail in reference . We plot both the excess noise introduced by the measurement, and the average final purity resulting from the measurement as a function of $`\theta `$ in Fig. 1. For a fixed measurement strength one therefore has the choice between choosing a measurement to minimize the noise, and consequently obtain better control of the system (in that the system will fluctuate less around the desired value), or obtain a more accurate knowledge of the system at the expense of increased noise. Which is most desirable may well depend upon the current state of knowledge. For example, if the state of the system is poorly known, perhaps early on in the control process, then it may prove desirable to obtain information more quickly, at the expense of introducing extra noise, since the large uncertainty will be the major factor in reducing the effectiveness of the feedback. However, once the observerโs knowledge is sufficiently sharp, it may prove more effective to reduce the noise at the expense of some added uncertainty. In Section VII we will present simulations to show how this information trade-off affects the performance of feedback control in a two-state system.
## VI Optimal Hamiltonian Feedback
Optimizing the Hamiltonian feedback for each feedback step consists in finding the feedback Hamiltonian which maximizes the fidelity with the target state, given the initial state of the system, which is fairly straightforward. Before we consider optimizing an infinitesimal Hamiltonian, which is necessary for continuous feedback, let us find the optimal unitary transformation. This result would be useful if a finite time were available for the feedback step, and feedback strength was not an important constraint, so that a โrotationโ through any angle could be performed by the feedback Hamiltonian.
Denoting the state of the system at the beginning of the feedback step as $`\rho `$, the fidelity with respect to the target state at the end of the feedback step is given by
$$F(\rho _\text{f},\sigma )=\text{Tr}[\sqrt{\rho ^{1/2}\sigma \rho ^{1/2}}],$$
(33)
where $`\sigma =|\psi _\text{T}\psi _\text{T}|`$ is the target state, and $`U`$ is the unitary transformation constituting the feedback. We wish to find $`U`$ to maximize $`F(\rho _\text{f},\sigma )`$. Observing first that, for arbitrary $`A`$ and unitary $`V`$,
$`\underset{V}{\mathrm{max}}|\text{Tr}[AV]|`$ $`=`$ $`\underset{V}{\mathrm{max}}|\text{Tr}[\sqrt{A^{}A}V^{}V]|`$ (34)
$`=`$ $`\underset{V}{\mathrm{max}}|{\displaystyle \underset{j}{}}\sigma _j(A)e^{i\theta _j}|`$ (35)
$`=`$ $`\text{Tr}[\sqrt{A^{}A}]`$ (36)
where we have used the polar decomposition theorem for $`A`$ ($`A=\sqrt{A^{}A}V^{}`$), and the $`\sigma _j(A)`$ are the eigenvalues of $`\sqrt{A^{}A}`$. Setting $`A=\rho _\text{f}^{1/2}\sigma ^{1/2}`$, this gives
$`F(\rho _\text{f},\sigma )`$ $`=`$ $`\underset{V}{\mathrm{max}}|\text{Tr}[(U\rho U^{})^{1/2}\sigma ^{1/2}V]|`$ (37)
$`=`$ $`\underset{V}{\mathrm{max}}|\text{Tr}[U\rho ^{1/2}U^{}\sigma ^{1/2}V]|`$ (38)
$``$ $`{\displaystyle \underset{j}{}}\lambda _j(\rho )^{1/2}\lambda _j(\sigma )^{1/2}`$ (39)
where the final inequality uses the result by von Neumann . In the last line $`\lambda _j(\rho )`$ and $`\lambda _j(\sigma )`$ are the eigenvalues of $`\rho `$ and $`\sigma `$ respectively, ordered such that the largest eigenvalue of $`\rho `$ multiplies the largest eigenvalue of $`\sigma `$, the second largest the second largest, and so on down to the smallest eigenvalue of both states. Now we need merely realize that we can achieve the upper bound by choosing $`U`$ so as to diagonalize $`\rho `$ in the basis of $`\sigma `$, re-ordering the basis states such that the largest eigenvalue of $`\rho `$ is attached to the eigenstate of $`\sigma `$ with the largest eigenvalue. Writing the eigenstates of $`\sigma `$ as $`|\sigma _j`$ (with eigenvalues ordered by size), and those of $`\rho `$ as $`|\rho _j`$ (similarly ordered), then the explicit construction for the optimal $`U`$ is
$$U=\underset{j}{}|\sigma _j\rho _j|.$$
(40)
For continuous observation, each feedback step acts only for an infinitesimal time, $`dt`$. In this case the final state, $`\rho _\text{f}`$, is given by
$$\rho _\text{f}=\rho i[H,\rho ]\mathrm{\Delta }t\frac{1}{2}[H,[H,\rho ]](\mathrm{\Delta }t)^2+\mathrm{},$$
(41)
where $`H`$ is the feedback Hamiltonian. The fidelity of the final state with the target state is therefore
$`\psi _\text{T}|\rho _\text{f}|\psi _\text{T}`$ $`=`$ $`\psi _\text{T}|\rho |\psi _\text{T}`$ (42)
$``$ $`i\psi _\text{T}|[H,\rho ]|\psi _\text{T}\mathrm{\Delta }t`$ (43)
$``$ $`\frac{1}{2}\psi _\text{T}|[H,[H,\rho ]]|\psi _\text{T}(\mathrm{\Delta }t)^2+\mathrm{}`$ (44)
The first term is fixed, so to optimize the fidelity in the infinitesimal time step we should maximize the coefficient of $`\mathrm{\Delta }t`$, being the dominant term. If the target state commutes with $`\rho `$, then this term vanishes for all $`H`$, so that we cannot choose a Hamiltonian that will cause an increase in the fidelity that is first order in time. If this situation occurs only for vanishingly small times, then it will make effectively no difference to the feedback performance. However, in those special situations in which the intrinsic dynamics preserves the commutivity of $`\rho (t)`$ and the target state, it can be important to choose a Hamiltonian which maximizes the term which is second order in time. One should note, however, that if one has freedom to choose the measurement basis, one can always choose a basis which disturbs the state so as to break the commutivity of $`\rho (t)`$ with the target state, eliminating the need to consider the second order term in the time evolution.
The maximization must be performed under a reasonable constraint on the eigenvalues of $`H`$ (i.e. a constraint that captures the concept of a limitation on the strength of feedback). A number of suitable constraints are possible, such as a restriction on the maximum eigenvalue of $`H`$, the sum of the norms of the eigenvalues, the sum of the squares of the eigenvalues, etc. Here we choose to use the last of these constraints, namely
$$\underset{n}{}\lambda _n(H)^2\mu .$$
(45)
To maximize the coefficient of $`\mathrm{\Delta }t`$ in Eq.(44), we first note that it may be written as the operator inner product $`\text{Tr}[HA]`$, where
$`A`$ $`=`$ $`i|\psi _\text{T}v|i|v\psi _\text{T}|,`$ (46)
$`|v`$ $`=`$ $`\rho |\psi _\text{T}`$ (47)
The maximum of the inner product, under the condition that the norms of the operators are constrained, occurs when the operators are aligned: $`H=cA`$, where $`c`$ is in general a complex number, but real in this case to preserve the Hermiticity of $`H`$. With this inner product the norm of $`H`$ is
$$\text{Tr}[H^2]=\lambda _n(H)^2.$$
(48)
Naturally, we take the maximum value allowed under the constraint, setting $`\text{Tr}[H^2]=\mu `$. This fixes the magnitude of the proportionality constant $`c`$, and results in the following explicit construction for the Hamiltonian
$$H=i\chi [|\psi _\text{T}\psi _\text{T}|,\rho ].$$
(49)
where
$$\chi =\sqrt{\frac{\mu }{ab^2}},$$
(50)
with
$`a`$ $`=`$ $`\psi _\text{T}|\rho ^2|\psi _\text{T}`$ (51)
$`b`$ $`=`$ $`\psi _\text{T}|\rho |\psi _\text{T}`$ (52)
It now remains to maximize the coefficient of $`(dt)^2`$ in Eq.(44). Recall this is only required under the condition that the first term is zero, which implies that $`\rho |\psi _\text{T}=\lambda _\text{T}|\psi _\text{T}`$. In this case the expression for the coefficient may be written as
$$\psi _\text{T}|H(\rho \lambda _\text{T}I)H|\psi _\text{T}.$$
(53)
Now, denoting the eigenvalues of $`\rho `$ by $`\lambda _n`$ (ordered in decreasing order), the eigenstates by $`|n`$, and denoting the target state as the eigenvector with $`n=M`$, the above expression becomes
$$\psi _\text{T}|H\rho H|\psi _\text{T}=\underset{n=1}{\overset{N}{}}(\lambda _n\lambda _M)|n|H|M|^2.$$
(54)
Now, the constraint on the Hamiltonian may be written
$`\text{Tr}[H^2]`$ $`=`$ $`\text{Tr}[{\displaystyle \underset{n}{}}|nn|H{\displaystyle \underset{m}{}}|mm|H]`$ (55)
$`=`$ $`{\displaystyle \underset{nm}{}}|n|H|m|^2=\mu ,`$ (56)
being a constraint on the sum of the square magnitudes of the elements of $`H`$. Only the subset $`|n|H|M|^2`$ of these, with $`nM`$, contribute to the expression to be maximized. To obtain the maximum value for the expression, we must therefore set all the elements of $`H`$ which do not contribute to it to zero, this allowing the contributing elements to be as large as possible. The constraint then becomes
$$\underset{nM}{\overset{N}{}}|n|H|M|^2=\mu /2,$$
(57)
where the factor of one half is enforced by the Hermiticity of $`H`$. The expression can now be seen as an average of the eigenvalues of $`\rho `$ over the โdistributionโ $`P(n)=|n|H|M|^2`$, which is normalized to $`\mu `$ by the constraint. The maximum value is therefore achieved when all the weight of the distribution is placed on the term with the largest eigenvalue. The solution is therefore
$$|1|H|M|^2=|M|H|1|^2=\mu /2,$$
(58)
with all other elements zero. The explicit construction for the optimal $`H`$ being
$$H=\sqrt{\mu /2}(|1\psi _\text{T}|+|\psi _\text{T}1|).$$
(59)
Note that this assumes that the target state is orthogonal to $`|1`$, being the eigenvector with the largest eigenvalue. If the $`|1`$ is the target state, then there exists no Hamiltonian evolution which will increase the fidelity, since the fidelity is the maximum it can be given the current purity of $`\rho `$. In that case we are free to set $`H=0`$ for that time step.
It is worth noting that since the magnitude of the feedback Hamiltonian and the strength of the continuous observation are uniformly bounded, the evolution of the system is continuous. Given this, since the Feedback Hamiltonian is a continuous function of the system state, it is intuitively clear that the feedback algorithm is well-defined (and continuous) in the continuum limit for almost all sample paths.
We now have a feedback algorithm which can be used in conjunction with a measurement strategy for feedback control. In the next section we will implement such a strategy for the control of a two-state system.
## VII Feedback Control of a Two-State System
In the previous sections we have considered the measurement and Hamiltonian feedback parts of the control problem separately. This resulted in a straightforward choice for a Hamiltonian feedback algorithm, but did not result in a clear choice for the measurement strategy. This was because we were able to identify in the measurement process a trade-off between information and disturbance. Because of this, the optimal measurement strategy for a given application is likely to depend upon the relative strengths available for the measurement and feedback. For example, if the feedback Hamiltonian is relatively strong, then it is likely that it will be able to effectively counter the disturbance introduced by the measurement, and therefore the measured observable should be chosen to provide maximal information and the expense of maximal disturbance. When this is not the case, the most desirable measurement is likely to be that which introduces less disturbance at the expense of providing reduced information.
To examine the performance of a feedback control algorithm we must run the algorithm many times in order to obtain the average behavior. This is computationally very expensive, and so we use massively parallel super-computers which are ideal for this task. The results we present here are obtained by averaging one thousand realizations of the control algorithm.
To provide a simple example of feedback control we consider a spin-half system precessing in a magnetic field aligned along the z-axis. In the absence of any noise, a spin aligned originally along the x-axis would rotate at a constant angular velocity around the z-axis, and we take this to be the desired (target) behavior. To provide the control problem, we subject the spin to noise which dephases it around the z-axis (this could arise from fluctuations in the magnetic field). The master equation describing the free (but noisy) evolution of the spin is thus given by
$$\dot{\rho }=i\mathrm{}\omega [\sigma _z,\rho ]\beta [\sigma _z,[\sigma _z,\rho ]],$$
(60)
where $`\omega `$ is the precession frequency in the magnetic field and $`\beta `$ is the strength of the dephasing noise. To implement feedback control we allow the observer (who is also naturally the controller) to measure the spin along an arbitrary spin direction $`๐ฏ(t)`$, with measurement constant $`k`$, and apply a feedback Hamiltonian, $`H_{\text{fb}}(t)`$, obtained using the algorithm presented in the previous section. The full evolution of the controllersโ state of knowledge, including the measurement and feedback, is therefore
$`d\rho `$ $`=`$ $`i\mathrm{}[\sigma _z+H_{\text{fb}}(t),\rho ]dt\beta [\sigma _z,[\sigma _z,\rho ]]dt`$ (63)
$`k[\sigma _{๐ฏ(t)},[\sigma _{๐ฏ(t)},\rho ]]dt`$
$`+\sqrt{2k}(\sigma _{๐ฏ(t)}\rho +\rho \sigma _{๐ฏ(t)}2\text{Tr}[\sigma _{๐ฏ(t)}\rho ]\rho )dW`$
We now simulate the dynamics resulting from the feedback control loop for different values of $`\theta `$, being the angle between the eigenbasis of the instantaneous system density matrix and the instantaneous measured observable, as discussed in Section V. For these simulations the strength of the magnetic field is such that $`\mathrm{}\omega =\pi `$, so that the spin rotates once in a time interval $`t=1`$. The noise strength is $`\beta =0.4`$, and the parameters for the control loop are $`k=2`$ and feedback strength $`\mu =10`$. We start the system in a pure state with the spin pointing along the x-direction, and evolve the controlled dynamics for a duration of $`t=2`$ (the purity and fidelity settle down to their steady state behavior by approximately $`t=0.8`$). Averaging the fidelity and purity over the full length of the run, for different values of $`\theta `$ we obtain figure 2. Examining the dependence of the purity on $`\theta `$, we find what we expect from the discussion in Section V. That is, the average purity of the system increases with $`\theta `$, achieving a maximum at $`\theta =\pi /2`$. This reflects the fact that, on average, measurements with a larger value of $`\theta `$ extract information from the system at a faster rate.
The behavior of the fidelity, in this case, is similar to that of the purity. As $`\theta `$ increases, the feedback is sufficient to ensure that even though we can expect the noise to increase with $`\theta `$, the increase in purity has more of an effect on the fidelity than the noise. The result is that, with these resources, it is best to choose $`\theta =\pi /2`$ (so as to measure in a basis maximally different from that which diagonalises the density matrix). However, from our previous analysis of the trade-off between information and disturbance, we cannot always expect this to be the case.
## VIII Conclusion
In this paper we have considered the problem of controlling a quantum system in real time using feedback conditioned on information obtained by continuous observation. The question of how to effect the best control given the system dynamics (including environmental noise), and constraints on available resources is highly non-trivial. Here we considered a simplified problem in which we examine the measurement process and the resulting Hamiltonian feedback separately. Our purpose was both to examine what concepts are motivated by the feedback control problem, and to explore the question of optimization in this simplified problem. A concept which arises immediately in considering feedback control is the strength of a measurement. This strength quantifies the amount of information which a measurement provides. Previous definitions of the information provided by quantum measurements have focussed on information regarding the state prior to the measurement. Here we have argued that it is the information regarding the state resulting from the measurement which is relevant to quantum feedback control, and introduced a concept of measurement strength accordingly.
Since measurements disturb quantum systems, it is important to understand how this relates to feedback control. We showed how it is possible to quantify the concept of the noise introduced by measurements in a way that is relevant to feedback control. One finds that while classical measurements do not introduce noise, quantum measurements do, in general, although it is possible, at least in principle, to make continuous quantum measurements that are noise free.
Having arrived at precise concepts of information and disturbance, we examined the special case of continuous measurements performed on a two-state system, and found that maximization of information, and minimization of noise were mutually exclusive goals, implying the existence of an information-disturbance trade-off in quantum feedback control. This highlights the complexity of the control problem.
We also considered the Hamiltonian feedback part of the control process. Defining the cost function as the fidelity with a target state, and the feedback strength as the norm of the Hamiltonian, we were able to obtain the Hamiltonian generating the optimal instantaneous feedback.
Here we explicitly consider control realized by choosing dynamics conditional upon a measurement process. Because of this one can refer to this technique as using a classical controller, since it works by taking a classical process (the measurement record) and altering the system Hamiltonian accordingly, all of which can be achieved by a classical system. It is therefore worth noting that, so long as we are considering the dynamics of the controlled system alone to be the important quantity, this is equivalent to control which is realized by connecting the system, via an interaction Hamiltonian, to another quantum system, where this second system is large enough to be treated as a bath . In general, using a second quantum system in this fashion may be referred to as using a quantum controller. When the quantum controller is finite dimensional and restricted in its dynamical response time, one can expect the performance of classical and quantum controllers to be somewhat different, and this is an interesting area for future work.
The question of how to best design feedback strategies to control noisy quantum systems is a complex one. However, the study of this problem will help us to understand better how quantum measurement may be exploited in the manipulation of quantum systems, and as quantum technology advances we can expect that this question will become increasingly important in practical applications.
## Acknowledgments
The authors would like to thank Howard Barnum, Tanmoy Bhattacharya, Chris Fuchs and Salman Habib for helpful discussions. KJ would like to thank Professor Sze Tan for hospitality during a visit to the University of Auckland where part of this work was carried out. This research was performed in part using the resources located at the Advanced Computing Laboratory of Los Alamos National Laboratory.
|
warning/0006/astro-ph0006300.html
|
ar5iv
|
text
|
# Dark energy and the Boomerang data
## I Introduction
In the last few years, a new reference model of structure formation has emerged, one in which 70% or so of the total matter content of the universe is in a form of dark energy. The main evidence for such a component is the supernovae Ia observations of Ref. and , which can be explained by assuming an accelerated expansion due either to a cosmological constant or to a new matter component with equation of state
$$p=(w1)\rho $$
(1)
with $`w(0,0.6)`$ . This new component can also be modeled as a light scalar field with a potential that allows for a potential-dominated epoch as, e.g., an exponential potential or an inverse power law .
The evidence for the new component, sometimes denoted dark energy or โquintessenceโ , has been reinforced by the recent Boomerang CMB observations that show a preference for a flat universe . In fact, since a CDM density of $`\mathrm{\Omega }_c(0.2,0.4)`$ is in agreement with a host of independent observations, from cluster abundance to cluster X-ray temperature, lensing, velocity fields etc., we can conclude that a conspicuous fraction of the total energy density has to be unclustered (or weakly clustered) and with negative pressure , although the direct constraint on the amount of dark energy from CMB alone is weak
The nature of this extra component is so far completely unknown. The simplest explanation, the cosmological constant, runs against the argument that, from a dimensional point of view, its value should be more than a hundred orders of magnitude smaller than expected, and tuned with astonishing precision in order to become dominant just today. The hypothesis of some fundamental scalar field has at least the advantage that one may hope to build a theory that explains the coincidence as a consequence of some fundamental principle. For instance, the inclusion of a coupling between dark matter and dark energy, as will be done in this paper, might explain why the two energy densities are comparable. Another motivation for considering a scalar field lies in the fact that it is premature to be too specific about the dark energy properties: a general scalar field includes as a limiting case the cosmological constant, but allows also to investigate less extreme effective equations of state.
In the same spirit of generality, I investigate the effect of an additional degree of freedom, represented by the coupling of the scalar field to ordinary matter. Such a coupling has been proposed and investigated in refs. , and is equivalent, up to a conformal rescaling, to the classical Brans-Dicke coupling to gravity (for gravity coupling in the context of quintessence see refs. ). If the dark energy is coupled to dark matter alone (and not to baryons), the present constraints on the coupling are rather weak ; as a consequence, in it was shown that the effects on the CMB turn out to be at the level of detectability already with the present data set.
Aim of this paper is to use the most recent CMB data, the Boomerang power spectrum , along with the COBE data (as reduced in ), to further constrain the coupled dark energy model. As particular cases, we will derive constraints on the pure cosmological constant model and on the uncoupled dark energy. A comparison of Boomerang with a different model of uncoupled dark energy is in .
In all the calculations of this paper a flat universe is assumed. This is of course a severe limitation, but the number of free parameters is already large enough to be at the limits of computational capabilities. Moreover, the hypothesis of quintessence has always been formulated in the context of flat models, in order to be consistent with the inflationary expectations.
While the work was almost completed the results of the Maxima experiment have been published. They seem to confirm the results of Boomerang, but have not been included in the present paper.
## II Coupled dark energy
The model of coupled quintessence or dark energy has already been studied in detail in . Here I limit myself to a summary of its properties.
Consider three components, a scalar field $`\varphi `$ , baryons and CDM, described by the energy-momentum tensors $`T_{\mu \nu (\varphi )},`$ $`T_{\mu \nu (b)}`$and $`T_{\mu \nu (c)}`$, respectively. General covariance requires the conservation of their sum, so that it is possible to consider a coupling such that
$`T_{\nu (\varphi );\mu }^\mu `$ $`=`$ $`\left(C_cT_{(c)}+C_bT_{(b)}\right)\varphi _{;\nu },`$ (2)
$`T_{\nu (c);\mu }^\mu `$ $`=`$ $`C_cT_{(c)}\varphi _{;\nu }.`$ (3)
$`T_{\nu (b);\mu }^\mu `$ $`=`$ $`C_bT_{(b)}\varphi _{;\nu }.`$ (4)
where $`C_{c,b}`$ are the coupling constants. This particular coupling is indeed obtained by conformally transforming a non-minimally coupled gravity theory or, following , by the Lagrangian (in units $`G=c=1`$)
$$L_{tot}=\frac{R}{2\kappa ^2}+\frac{1}{2}\varphi _{;\mu }\varphi ^{;\mu }U(\varphi )+L_c(e^{2C_c\varphi }g_{\mu \nu })+L_b(e^{2C_b\varphi }g_{\mu \nu })$$
(5)
where $`\kappa ^2=8\pi `$ and where $`L_{c,b}`$ denote the Lagrangian of the dark matter and baryonic fields. Notice that our constant $`C_c`$ corresponds to the invisible coupling constant $`\beta _I`$ of . The radiation field (subscript $`\gamma `$) remains uncoupled, since $`T_{(\gamma )}=0.`$ We derive the background equations in the flat conformal FRW metric $`ds^2=a^2(d\tau ^2+\delta _{ij}dx^idx^j)`$ assuming the exponential potential
$$U=Ae^{s\varphi }$$
(6)
as proposed e.g. in . As anticipated, we will couple the dark energy scalar field to the dark matter only, putting $`C_b=0`$. We call this choice dark-dark coupling. Generalizing we introduce the following variables (putting $`H=\dot{a}/a`$)
$$x=\frac{\kappa }{H}\frac{\dot{\varphi }}{\sqrt{6}},y=\frac{\kappa a}{H}\sqrt{\frac{U}{3}},z=\frac{\kappa a}{H}\sqrt{\frac{\rho _\gamma }{3},}v=\frac{\kappa a}{H}\sqrt{\frac{\rho _b}{3},}$$
(7)
Adopting the $`e`$-folding time $`\alpha =\mathrm{log}a`$ we can write the field, baryon and radiation conservation equation as a system in the four variables $`x,y,z,v`$ that depends on the parameters $`\mu ,\beta `$
$`x^{}`$ $`=`$ $`\left({\displaystyle \frac{z^{}}{z}}1\right)x\mu y^2+\beta (1x^2y^2z^2v^2),`$ (8)
$`y^{}`$ $`=`$ $`\mu xy+y\left(2+{\displaystyle \frac{z^{}}{z}}\right),`$ (9)
$`z^{}`$ $`=`$ $`{\displaystyle \frac{z}{2}}\left(13x^2+3y^2z^2\right),`$ (10)
$`v^{}`$ $`=`$ $`{\displaystyle \frac{v}{2}}\left(3x^2+3y^2z^2\right)`$ (11)
where the prime denotes derivation with respect to $`\alpha `$, and where $`\beta =C_c\sqrt{\frac{3}{2\kappa ^2}},\mu =s\sqrt{\frac{3}{2\kappa ^2}}.`$ The coupling constant $`\beta `$ can therefore be regarded as the ratio of the dark energy coupling strength to the gravitational strength. The dark-dark coupling $`\beta `$ is unconstrained by local experiments or by $`\dot{G}/G`$ measurements . Constraints can be derived only by global effects on the cosmological dynamics: for instance, from the requirement of a universe older than 10 Gyr Ref. found
$$|\beta |<1$$
(12)
In the following we show that Boomerang puts a much more stringent constraint on the dark-dark coupling.
Neglecting the baryons, the system has a transient attractor (properly speaking, a saddle point) at the point $`x_0=2\beta /3,y_0=0,z_0=0`$ on which the scale factor expands as $`at^{6/(9+4\beta ^2)}`$ (here $`t`$ is the ordinary time defined as $`d\tau =a(t)dt`$), that is, slower than in a ordinary MDE. During this stage, first found in Ref. , and denoted $`\varphi `$MDE in , the scalar field kinetic energy and the dark matter have a constant density ratio. When the universe leaves the saddle, it reaches a global attractor if $`\mu <3`$, at the point $`x_0=\mu /3,y_0=(1\mu ^2/9)^{1/2},z_0=0`$ and the universe expansion follows the law $`at^{3/\mu ^2}`$ mimicking a perfect fluid with equation of state $`p=(w_\varphi 1)\rho `$, with $`w_\varphi =2\mu ^2/9`$ . The expansion is accelerated only if $`\mu <\sqrt{3}`$. As long as the baryon component is small with respect to the CDM, the phase space of the system remains qualitatively the same.
It can be noticed that the $`\varphi `$MDE depends only on $`\beta `$, while the subsequent $`\varphi `$-dominated epoch depends only on $`\mu `$. Therefore, as shown in , the possibility to set constraints on the coupling $`\beta `$ depends on the existence of the $`\varphi `$MDE. This epoch is actually quite more general than appears from the discussion above: it exists infact for any potential such that $`U(\varphi )`$ and $`dU/d\varphi `$ go asymptotically to zero, as for instance when $`U`$ is an inverse power law.
The coupled quintessence with exponential potential contains the case of pure cosmological constant ($`\mu =0,\beta =0`$), of uncoupled quintessence ($`\beta =0`$) and is asymptotically equivalent to a perfect fluid with a constant equation of state $`w_\varphi =2\mu ^2/9`$. The model is conformally equivalent to a non-minimal gravity theory with Lagrangian term $`\frac{1}{2}\xi \psi ^2R`$ and potential $`V(\psi )=\lambda \psi ^n`$ with $`\xi =2\beta ^2/3`$ and $`n=4+\mu /\beta `$ . Therefore, a bound on $`\beta `$ amounts to constraining the non-minimal gravity coupling $`\xi `$. Notice that in the case we study here, in which the dark energy couples differently to baryons and to dark matter, there are two conformally related metrics in the Jordan frame (the frame in which gravity is coupled to the scalar field): the constant $`\xi `$ couples the scalar field to the Ricci scalar expressed in the metric in which the dark matter, rather than the baryons, follow the geodetics (see ).
The perturbation equations, derived in , are integrated by the use of a purposedly modified CMBFAST code . In addition to the scalar field, baryons, CDM and radiation, the code includes also massless neutrinos. I choose adiabatic initial conditions, as suggested by inflation. The initial conditions for the background equations are found for each set of parameters by trial and error so as to give today $`\mathrm{\Omega }_b,\mathrm{\Omega }_c`$ and $`H_0`$ as requested. This procedure is so time consuming with respect to the code without coupled scalar field that the assumption of flat space and a further reduction in the parameter space explained below turned out to be necessary.
## III Likelihood analysis
Our theoretical model depends on two scalar field parameters and four cosmological parameters:
$$\beta ,\mu ,n_s,h,\mathrm{\Omega }_b,\mathrm{\Omega }_c$$
(13)
The remaining input parameters requested by the CMBFAST code are set as follows: $`T_{cmb}=2.726K,`$ $`Y_{He}=0.24,N_\nu =3.04,\tau _c=0.`$ The latter quantity specifies the optical depth to Thomson scattering, and is a measure of reionization. In the analysis of this parameter was also included in the general likelihood and, in the flat case, was found to be compatible with zero at slightly more than 1$`\sigma `$ . Moreover, in ref. it is found that fixing $`\tau =0`$ has only a minor effect on the other parameters. Therefore here, to further reduce the parameter space, I assume $`\tau _c`$ to vanish. Two other approximations with respect to have been necessary: first, I did not include the cross-correlation between bandpowers because it is not available. Second, an offset log-normal approximation to the band-power likelihood has been advocated by and adopted by , but the $`x`$-quantity necessary for its evaluation is not available. Since the offset log-normal reduces to a log-normal in the limit of small noise I evaluated the log-normal likelihood
$$2\mathrm{log}L(\alpha _j)=\underset{i}{}\frac{\left[Z_{\mathrm{},t}(\mathrm{}_i;\alpha _j)Z_{\mathrm{},d}(\mathrm{}_i)\right]^2}{\sigma _{\mathrm{}}^2}$$
(14)
where $`Z_{\mathrm{}}\mathrm{log}\widehat{C}_{\mathrm{}}`$, the subscripts $`t`$ and $`d`$ refer to the theoretical quantity and to the real data, $`\widehat{C}_{\mathrm{}}`$ are the spectra binned over some interval of multipoles centered on $`\mathrm{}_i`$, $`\sigma _{\mathrm{}}^2`$ are the experimental errors on $`Z_{\mathrm{},d}`$, and the parameters are denoted collectively as $`\alpha _j`$. A 10% calibration error is added to the experimental errors . I also evaluated for comparison a likelihood gaussian in the variables $`\widehat{C}_{\mathrm{}}`$, and found that the results do not change appreciably up to 2$`\sigma `$ from the peak.
Among the parameters that refer to the scalar field, we have already shown in that $`\mu `$ has a negligible effect on the background solution at $`z1`$, since the equivalence time does not depend on it and, although $`\mu `$ sets the speed of the present accelerated expansion, this has only a minor effect on the perturbation spectrum at decoupling. This will be confirmed below. Also, in we have shown that the dynamics of the system is insensitive to the sign of $`\beta `$, since both the $`\varphi `$MDE and the final accelerated epoch do not depend on it. We will consider only $`\beta 0`$.
In order to compare with the Boomerang analysis I assume uniform priors as in , with the parameters confined in the range $`\beta (0,0.16),`$ $`\mu (0,2.1),`$ $`n_s(0.7,1.3),`$ $`h(0.45,0.9),`$ $`\mathrm{\Omega }_b(0.01,0.2),`$ $`\mathrm{\Omega }_c(0.1,0.9),`$ with the further weak big-bang nucleosynthesis (BBN) constraint $`\mathrm{\Omega }_bh^2<0.05`$ (and of course the condition $`\mathrm{\Omega }_c+\mathrm{\Omega }_b1`$) . I evaluate the likelihood also using the most up-to-date BBN limit , assuming a gaussian prior for $`\mathrm{\Omega }_bh^2`$ with mean $`0.019`$ and 1$`\sigma `$ error $`0.002`$ (I refer to this as the strong BBN prior). The range of $`\mu `$ includes all the values for which there is acceleration at the present. A further parameter, the logarithm of the absolute normalization of the $`C_{\mathrm{}}`$ spectrum, is integrated out analytically with uniform prior. The same age constraints ($`>10`$ Gyr) used in is adopted here. Note that here we do not have the problem of near-degeneracy of parameters that is considered in ref. . In fact, the degeneracy that exists for those values of $`\mathrm{\Omega }_c,\mathrm{\Omega }_b,h,\mu `$ and $`\beta `$ that give an identical angular-diameter distance to the decoupling surface is completely removed by the combined constraints of flatness, BBN and the allowed range of $`h`$ and $`\mu `$. We also checked that the limits on $`n_s`$ and $`\beta `$ are broad enough to contain the bulk of likelihood.
A grid of $`10,000`$ multipole CMB spectra $`C_{\mathrm{}}`$ is used as a database over which I interpolate to produce the likelihood function. Since both the COBE and the Boomerang data are in fact binned over intervals of multipoles, I average in the same bandpowers the theoretical spectra for a correct comparison. Three cases will be distinguished: pure $`\mathrm{\Lambda }`$ ($`\mu =\beta =0`$); dark energy ( $`\beta =0`$); coupled dark energy. In Fig. 1 I report examples of multipole spectra obtained varying $`\mu `$ and $`\beta `$ and fixing the other parameters: $`\{n_s,\mathrm{\Omega }_b,\mathrm{\Omega }_c,h\}=\{1,0.05,0.3,0.7\}`$. The quantity plotted is actually $`(\mathrm{}(\mathrm{}+1)C_{\mathrm{}}/2\pi )^{1/2}\mu `$K as customary. The strong decrease in amplitude of the acoustic peaks as $`\beta `$ gets larger than 0.1 depends on the fact that for these values the onset of the $`\varphi `$MDE occurs before the decoupling. During the $`\varphi `$MDE the fluctuations smaller than the horizon grow less than during MDE, so that the fluctuations on the acoustic scales are depressed relative to the larger scales. Moreover, the addition of the integrated Sachs-Wolfe effect at small multipoles decreases the normalization of the intrinsic fluctuations at decoupling .
In Fig. 2 ( panels $`ad`$, solid lines) I plot the likelihood function for the parameters $`n_s,h,`$ $`\mathrm{\Omega }_b,\mathrm{\Omega }_c`$ (marginalizing over all the other parameters) in the case of pure $`\mathrm{\Lambda }`$ and weak BBN prior. They are reasonably well in agreement with the analysis of (their case P10). The means and variances for the pure $`\mathrm{\Lambda }`$ model are
$$n_s=0.96\pm 0.06,h=0.73\pm 0.09,\mathrm{\Omega }_b=0.056\pm 0.016,\mathrm{\Omega }_c=0.5\pm 0.2,$$
(15)
while the best estimates (peaks) are $`n_s=0.975,h=0.7,\mathrm{\Omega }_b=0.05,\mathrm{\Omega }_c=0.3`$ in agreement with (notice that Ref. quotes the likelihood maximum for clear detections, e.g. for $`\mathrm{\Omega }_b`$ and $`n_s`$, and the mean in the other cases). For the case of uncoupled dark energy, I plot in Fig. 2 the likelihood functions in short dashed lines, plus in the panel $`e`$ the new parameter $`\mu `$. As expected, there is almost complete degeneracy in the direction of $`\mu `$, and the results of pure $`\mathrm{\Lambda }`$ remain very similar. If the Boomerang errorbars were reduced to one third, the likelihood would begin to show some preference for higher values of $`\mu `$, as shown in the same panel $`e`$ (dotted line). The problem of a high baryon content from Boomerang (see e.g. ref. ) is not alleviated by this model of dark energy.
The likelihood functions for the coupled dark energy case are plotted in Fig. 2 as long-dashed lines. Due to the degeneracy along $`\mu `$ we can simplify the analysis fixing this parameter to any value in the relevant range; we put $`\mu =0.25`$. It turns out that the previous results are quite robust also with respect to the coupling, except for a shift of $`n_s`$ toward smaller values ($`n_s=0.88\pm 0.06`$). In the panel $`f`$ I plot the coupling $`\beta `$ : at the 96.8% c.l. we obtain
$$|\beta |<0.1$$
(16)
Assuming the strong BBN prior we obtain the likelihood functions shown as dot-dashed lines in the same Fig. 2 (shown only for the coupled dark energy case for clarity; the other cases are similar). As expected, smaller $`\mathrm{\Omega }_b`$ and smaller $`h`$ than before are now acceptable. Also, $`\mathrm{\Omega }_c`$ moves to smaller values, in order not to decrease excessively the first peak. This effect, along with the decrease of $`n_s`$ in panel $`b`$, is observed also in when the strong BBN is imposed. The likelihood for the exponential slope $`\mu `$ (not shown) becomes quite flatter with the strong BBN. Interestingly, $`\beta `$ now peaks around 0.05, because this value raises somewhat the first peak, compensating for the low baryon content. The limit (16) remains valid.
## IV Conclusions
The most interesting conclusion that can be drawn from our analysis is that the coupling $`\beta `$ between dark matter and dark energy has to be smaller than 0.1, an order of magnitude better than previously , independently of the BBN constraint. Although I derived this limit in the particular case of an exponential potential, it extends to a much larger class of theories since the $`\varphi `$MDE is dominated by the kinetic, rather than potential, energy of the scalar field, as will be shown in another paper,
In terms of the coupling constant $`\xi `$ of non-minimal gravity theories we get roughly $`\xi <0.01`$ at the 96.8% c.l. .
The likelihood of the other cosmological parameters is robust with respect to the addition of the coupling.
The potential parameter $`\mu `$, which sets the effective equation of state of the dark energy at the present, is not well constrained by the present CMB, since its effect is very recent on the cosmological time scale.
It is of course possible to add further constraints from large-scale structure, supernovae Ia and from other CMB experiments. In a constraint similar to (16) was found from cluster abundance, for a particular choice of the other parameters. Future data, especially high-resolution, high-coverage experiments on CMB, have the potential to strenghten the limit by at least another order of magnitude.
|
warning/0006/math0006193.html
|
ar5iv
|
text
|
# Quantum periods โ I. Semi-infinite variations of Hodge structures.
## 1. Introduction
This paper is concerned with theory of periods of non-commutative varieties which arise naturally in the context of mirror symmetry.
We have noticed in \[B1\] that in order to understand mirror symmetry phenomena in dimensions higher then three one has to enlarge the class of functions provided by periods of complex manifolds. In this paper we introduce a new class which consists of periods associated with *non-commutative* deformations of complex manifolds. We show that this is precisely the class which is needed in order to generalize to dimensions higher then three the mirror symmetry relation between rational Gromov-Witten invariants of Calabi-Yau manifolds and periods associated with their mirror partners.
As another motivation for developing theory of quantum periods let us mention that such a theory should serve as one of guides for future theory of non-commutative varieties. It is interesting to make a somewhat ambitious parallel and to note that, historically, developement of integration theory of analytic functions (theory of elliptic integrals, in particular, and, more generally, of Abelian integrals) was one of the principal reasons for introduction of Riemann surfaces.
One may hope also that quantum periods should help to understand certain aspects of varieties from ordinary commutative world (see, for example, the local Torelli property of the period map from ยง6.3) and, in particular, especially those related with $`K`$theory.
We plan to write a series of papers concerning the theory of quantum periods and its applications. Next paper (\[B2\]) is devoted to applications to non-commutative geometry and homological mirror symmetry. In a separate paper (\[B3\]) we consider in details an oversimplified situation where periods of somewhat related type appear. There we deal with semi-infinite Hodge structure associated with mirror partners of projective spaces. The main result of \[B3\] is the equality expressing generating functions for total sets of rational Gromov-Witten invariants of $`^n`$in terms of periods of semi-infinite Hodge structures associated with their mirrors.
Let us say a few words on the plan of the paper. Let $`X`$ denotes a complex smooth manifold. We begin in ยง2 by recalling description of the moduli space $``$ of non-commutative deformations of $`X`$ via differential graded Lie algebra $`๐ค_A_{\mathrm{}}`$ of infinitesimal $`A_{\mathrm{}}`$symmetries of $`X`$. The moduli space $``$ is the quotient of graded scheme of zeroes of solutions to Maurer-Cartan equation in $`๐ค_A_{\mathrm{}}`$ with respect to gauge equivalences. In ยง3 we show that Gauss-Manin local system extends naturally over the moduli space $``$. This is based on the general observation (prop.3.2) that sheaves with flat connection over moduli space associated with differential graded Lie algebra $`๐ค`$ can be described via modules over $`๐ค[\xi ]/\xi ^2`$. In ยง4 we associate with moduli space $``$ semi-infinite analog of variations of Hodge structures. If $`\gamma (๐ค_A_{\mathrm{}}๐_R)^1`$ denotes a solution to Maurer-Cartan equation describing a non-commutative deformation of $`X`$ given by $`\varphi (R)`$ then the semi-infinite analog of Hodge filtration associated with $`\varphi `$ is the subspace $`L(\varphi )Gr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])(R)`$ which consists of elements of the form
$$\left[l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(\phi _0+\phi _1\mathrm{}+\phi _2\mathrm{}^2\mathrm{})\right]\text{where}\phi _i\mathrm{\Omega }^{}(X,)R$$
where $`i_\gamma `$ denotes contraction $`\mathrm{\Lambda }^{}T\mathrm{\Omega }^{}\mathrm{\Omega }^{}`$extended by wedge multiplication by $`(0,)`$forms and $`l_{\mathrm{}}`$ denotes rescaling operator $`\phi ^{p,q}\mathrm{}^{\frac{qp}{2}}\phi ^{p,q}`$ for $`\phi ^{p,q}\mathrm{\Omega }^{p,q}`$. In ยง5 we study variations of the subspaces $`L(\varphi )`$, $`\varphi (R)`$ and prove that they satisfy semi-infinite version of Griffiths transversality condition:
$$L\mathrm{}^1L$$
In ยง6 we explain the role played by periods of such semi-infinite Hodge structures in higher dimensional mirror symmetry. Similar to the three-dimensional case (\[COGP\]) the calculation of Gromov-Witten invariants of an $`n`$dimensional Calabi-Yau manifold $`Y`$ via quantum periods associated with mirror Calabi-Yau manifold $`X`$ is done in three steps. At the first step one specifies normalization of the periods. Assume that we are given a linear subspace $`SH[[\mathrm{}^1,\mathrm{}]]`$, $`\mathrm{}^1SS`$ transversal to subspace $`L(0)`$ and an element $`\psi _0L(0)`$. Such linear subspace $`S=L_W`$ and element $`\psi _0`$ are associated naturally with monodromy weight filtration $`W`$ of maximal unipotency cusp on moduli space of complex structures on $`X`$. Then there exists unique element $`\mathrm{\Psi }^W(t,\mathrm{})L(t)`$, $`t`$ satisfying the normalization condition $`\mathrm{\Psi }^W(t,\mathrm{})(L_W+\psi _0)`$. Second step is a choice of coordinates on the moduli space $``$ using the periods normalized in the first step. The coordinates are induced via map to linear space $`L_W/\mathrm{}^1L_W`$ defined as $`(\mathrm{\Psi }^W(t,\mathrm{})\psi _0)\text{mod}\mathrm{}^1L_W`$. In the last step one writes differential equation:
$$\frac{^2\mathrm{\Psi }^{W,\alpha }}{t_W^at_W^b}=\mathrm{}^1\underset{c}{}A_{ab}^c(t_W)\frac{\mathrm{\Psi }^{W,\alpha }}{t_W^c}$$
satisfied by quantum periods $`\mathrm{\Psi }^{W,\alpha }(t_W)=_{\mathrm{\Delta }_\alpha }\mathrm{\Psi }^W`$ where $`\{\mathrm{\Delta }_\alpha \}`$ is a basis in $`_kH_k(X,)`$ and $`\{t_W^a\}`$ are the coordinates from step two. Then $`A_{ab}^c(t_W)`$ should coincide with structure constants of (big) quantum cohomology of $`Y`$ and the coefficients of appropriate Taylor expansion of $`A_{ab}^c(t_W)`$ should give the set of all rational Gromov-Witten invariants of $`Y`$. We checked this in \[B1\] ยง6 for Calabi-Yau complete intersections varieties.
## 2. Moduli spaces of $`A_{\mathrm{}}`$deformations of complex manifolds.
It is explained here how to think about moduli space $``$ of non-commutative deformations of complex manifold $`X`$. We assume that the reader is familiar with the technique of deformation theory (see for example, \[K2\] ยง4, \[B1\] ยง2, and references therein). Idea of considering $`A_{\mathrm{}}`$deformations of $`D^bCoh(X)`$ was suggested in \[K1\]. The definition of the moduli space given below appeared in \[BK\].
We would like to stress that a satisfactory definition of objects that are parameterized by moduli space $``$ remains yet a mystery. The only thing which is known is that it should be some quantum varieties such that their equivalence classes are in one-to-one correspondence with equivalence classes of $`A_{\mathrm{}}`$deformations of bounded derived category of coherent sheaves on $`X`$. Remark that, as it follows easily from the definition of $``$ given below, our moduli space contains a closed subspace which parameterizes classes of equivalences of Poisson brackets on $`X`$. The space of equivalence classes of first-order deformations parameterized by $``$ is given by $`T_{[X]}=_{p,q}H^q(X,\mathrm{\Lambda }^pT_X)[pq]`$. The subspace of Poisson brackets corresponds on the level of first-order deformations to the subspace $`H^0(X,\mathrm{\Lambda }^2T_X)(_{p,q}H^q(X,\mathrm{\Lambda }^pT_X)[pq])`$.
### 2.1. Deformations of complex structure
Recall first the standard description of deformations of complex structures. Given complex manifold $`X`$ let $`J`$ denotes the corresponding complex structure on the underlying $`C^{\mathrm{}}`$manifold $`X_C^{\mathrm{}}`$. According to Kodaira-Spencer theory the deformations of $`J`$ are described by elements $`\rho \mathrm{\Omega }^{0,1}(X,T)`$ satisfying
(2.1)
$$\overline{}\rho +\frac{1}{2}[\rho ,\rho ]=0$$
Namely, to an element $`\rho `$ one can associate a deformation of the decomposition $`T_{}^{}=T^{}\overline{T}^{}`$of bundle of $``$valued $`1`$forms into the sum of sub-bundles corresponding to differentials of holomorphic and anti-holomorphic coordinates. The deformed sub-bundle $`T^{}`$ is given by the graph of the linear map corresponding to $`\rho \mathrm{\Gamma }(X,T\overline{T}^{})`$. Eq. (2.1) means that this sub-bundle defines formally integrable distribution. Such a distribution gives a new complex structure on $`X_C^{\mathrm{}}`$ by Newlander-Nirenberg theorem. Also, equivalent deformations are related via the action of the Lie algebra $`\mathrm{\Gamma }_C^{\mathrm{}}(X,T)`$. Notice that this description is formulated intrinsically in terms of of differential graded Lie algebra on $`_q\mathrm{\Omega }^{0,q}(X,T)[q]`$. We denote this differential graded Lie algebra via $`๐ค_{cs}`$.
### 2.2. $`A_{\mathrm{}}`$deformations
The $`A_{\mathrm{}}`$deformations of $`X`$ allow similar description using elements $`\gamma \mathrm{\Omega }^{0,}(X,\mathrm{\Lambda }^{}T)`$. Consider the differential graded Lie algebra
$$๐ค_A_{\mathrm{}}:=\underset{i}{}๐ค_A_{\mathrm{}}^i[i],๐ค_A_{\mathrm{}}^i=\underset{qp+1=i}{}\mathrm{\Omega }^{0,q}(X,\mathrm{\Lambda }^pT)$$
equipped with Dolbeault differential and the bracket which is given by Schouten-Nijenhuis bracket on $`\mathrm{\Lambda }^{}T`$ extended using cup-product of anti-holomorphic differentials. Consider the graded scheme of solutions to Maurer-Cartan equation
$$MC_A_{\mathrm{}}:=\{\overline{}\gamma +\frac{1}{2}[\gamma ,\gamma ]=0\}๐ค_A_{\mathrm{}}[1]$$
The standard approach consists of developing solutions to Maurer-Cartan in power series $`\gamma =_a\gamma _at^a+\frac{1}{2}_{a_1a_2}\gamma _{a_1a_2}t^{a_1}t^{a_2}+\mathrm{}`$. In this paper we leave aside the question of convergence of power series representing the solutions, which can be treated separately using the standard technique of Kuranishi spaces. Consequently, we will work on the level of formal power series and will be dealing with the completion of $`MC_A_{\mathrm{}}`$ at zero. In the sequel $`MC_A_{\mathrm{}}`$ will always denote the corresponding formal scheme. This scheme has points defined over arbitrary graded $``$Artin algebra<sup>1</sup><sup>1</sup>1Recall that any such algebra is isomorphic to $`[t_s]_{sS}/I`$, where $`S`$ is some finite set and $`I`$ is an ideal containing sufficiently large power of the maximal ideal $`t[t_s]`$. $`R`$:
$$MC_A_{\mathrm{}}(R)=\{\gamma |\overline{}\gamma +\frac{1}{2}[\gamma ,\gamma ]=0,\gamma (๐ค_A_{\mathrm{}}[1]๐_R)^0\}$$
where $`๐_R`$ denotes the unique maximal ideal in $`R`$. We will also consider points of $`MC_A_{\mathrm{}}`$ over projective limits of Artin algebras<sup>2</sup><sup>2</sup>2Here and below by Artin algebra we mean graded $``$Artin algebra if it does not specified otherwise. , for example, over algebra of formal power series on some finite-dimensional $``$graded vector space. Algebra Lie $`๐ค_A_{\mathrm{}}`$ acts on the scheme $`MC`$ via gauge transformations (again it should be understand as action in the category of formal graded schemes):
(2.2)
$$\alpha ๐ค_A_{\mathrm{}}\dot{\gamma }=\overline{}\alpha +[\gamma ,\alpha ]$$
For any Artin algebra $`R`$ the algebra Lie $`(๐ค_A_{\mathrm{}}๐_R)^0`$is nilpotent and the associated group $`\mathrm{exp}\left((๐ค_A_{\mathrm{}}๐_R)^0\right)`$ acts on $`MC_A_{\mathrm{}}(R)`$. Explicitly, the action of an element $`\phi =\mathrm{exp}\alpha \mathrm{exp}\left((๐ค_A_{\mathrm{}}๐_R)^0\right)`$ is given by
(2.3)
$$\gamma \gamma ^\phi :=\overline{}\phi \phi ^1+\phi \gamma \phi ^1=\underset{n=0}{\overset{\mathrm{}}{}}(\frac{(ad\alpha )^{n+1}\overline{}\alpha }{(n+1)!}+\frac{(ad\alpha )^n\gamma }{n!})$$
###### Definition.
The (formal) moduli space $``$ is the quotient of $`MC_A_{\mathrm{}}`$by the gauge transformations (2.3).
More precisely this should be understand as follows: $``$ is described via the functor $`Def_{๐ค_A_{\mathrm{}}}^{}`$: $``$graded Artin algebras $``$sets, which is given by
(2.4)
$$Def_{๐ค_A_{\mathrm{}}}^{}(R):=MC_A_{\mathrm{}}(R)/\mathrm{exp}\left((๐ค_A_{\mathrm{}}๐_R)^0\right)$$
Morally the value of this functor $`Def_{๐ค_A_{\mathrm{}}}^{}(R)`$ on Artin algebra $`R`$ is the set of morphisms $`(\text{Spec}R,0)`$. The tangent space to $``$ at the base point (more precisely the space of equivalence classes of first-order deformations) is equal to $`Def_{๐ค_A_{\mathrm{}}}^{}([t]/t^2)`$ which gives
$$T_{[X]}=_{p,q}H^q(X,\mathrm{\Lambda }^pT)[pq]$$
In the sequel we will denote via $`(R)`$ the set of โpointsโ of $``$ over $`R`$, i.e. set of equivalence classes of $`A_{\mathrm{}}`$deformations depending on parameters from $`R`$.
If one restricts oneself to solutions to Maurer-Cartan equation from $`MC_{cs}(R):=\{\gamma |\overline{}\gamma +\frac{1}{2}[\gamma ,\gamma ]=0,\gamma \mathrm{\Omega }^{0,1}(X,T)๐_{R^0}\}`$ then it corresponds to deformations of complex structure on $`X`$. We denote by $`^{cs}`$ the quotient of $`MC_{cs}`$ by gauge equivalences (2.2). It is easy to see that such equivalences are generated only by elements from $`๐ค_{cs}^0=\mathrm{\Gamma }_C^{\mathrm{}}(X,T)`$. For compact complex manifold $`X`$ the formal moduli space $`_{cs}`$ is isomorphic to the completion at the base point $`[X]`$ of Kuranishi space which is a mini-versal family of deformations of complex structure on $`X`$ (Goldman-Millson) (here we assumed for simplicity that $`H^0(X,T_X)=0`$, otherwise the statement is slightly more complicated).
The formality theorem from \[K2\] can be used to show that the moduli space $``$ parameterizes $`A_{\mathrm{}}`$deformations of the category $`D^bCoh(X)`$. The corresponding arguments are sketched in ยง3.3 of \[B1\].
There is a useful point of view on the quotients of type (2.4) (see \[K2\] ยง4, \[B1\] ยง2). Such quotients can be more generally associated with germs of $`Q`$manifolds (or dg-manifolds in different terminology). This is a germ of $``$graded manifold plus a degree one vector field $`Q`$ satisfying $`[Q,Q]=0`$. Any differential graded Lie algebra can be viewed as $`Q`$manifold. Namely, vector field
$$Q_๐ค(\gamma )=d\gamma +\frac{1}{2}[\gamma ,\gamma ]$$
on $`๐ค[1]`$ satisfies $`[Q_๐ค,Q_๐ค]=0`$. Then the moduli space described by $`๐ค`$ can be viewed as a kind of nonlinear cohomology associated with $`Q_๐ค`$. More precisely, it is realized as a quotient
(2.5)
$$_๐ค=(\text{zeros}\text{of}Q)/\text{distribution}\text{generated}\text{by}\text{vector}\text{fields}\text{of}\text{the}\text{form}[Q,])$$
of the type similar to (2.4). It is often useful to express statements about the quotient in $`Q`$equivariant terms on the underlying $`Q`$-manifold.
Let us add that it is natural to expect that for complex algebraic manifold $`X`$ the moduli space $``$ can be equipped with appropriate algebraic structure i.e. $``$ must be defined canonically on the level of (a kind of) dg-scheme.
## 3. Gauss-Manin local system over $``$.
This section is devoted to an analog of Gauss-Manin local system living over the moduli space of $`A_{\mathrm{}}`$deformations of $`X`$.
Sheaves over moduli spaces are described by modules over corresponding differential graded Lie algebras. The easiest way to see this is to use the geometric language of $`Q`$manifolds. Let $`๐ค`$ denotes differential graded Lie algebra describing moduli space $`_๐ค`$ as in (2.4). Given dg-module $`h`$ over $`๐ค`$ with its structure morphisms given by linear maps $`d_h:hh[1]`$, $`m:๐คhh`$ one can define $`Q_๐ค`$connection $`๐_{Q_๐ค}`$ on trivial bundle over $`๐ค[1]`$ with fibers equal to the linear space $`h`$. The covariant derivatives of constant sections are given by
(3.1)
$$(๐_{Q_๐ค}s)(\gamma ):=d_hs+m(\gamma s)\text{for}sh$$
Using the Leibnitz rule the covariant derivative $`๐_{Q_๐ค}s`$ can be defined for arbitrary section of the trivial bundle. The conditions imposed on linear maps $`(d_h,m)`$ in order that they define dg-module structure on $`h`$ can be reformulated simply as $`[๐_{Q_๐ค},๐_{Q_๐ค}]=0`$. It follows that for $`\gamma MC_๐ค(R)`$ one has twisted differential acting on the fiber over $`\gamma `$ giving rise to complex $`(hR,๐_{Q_๐ค}|_\gamma )`$.
###### Proposition 3.1.
Given a solution to Maurer Cartan equation $`\gamma MC_๐ค(R)`$ and an element of the group of gauge transformations $`\phi \mathrm{exp}\left((๐ค_A_{\mathrm{}}๐_R)^0\right)`$ the cohomology of complexes $`(hR,๐_{Q_๐ค}|_\gamma )`$ and $`(hR,๐_{Q_๐ค}|_{\gamma ^\phi })`$ are canonically isomorphic and the correspondence which associates to $`\gamma MC_๐ค(R)`$ the cohomology of complex $`(hR,๐_{Q_๐ค}|_\gamma )`$ defines a sheaf over the moduli space $`_๐ค`$.
###### Proof.
Morally this means that $`Q_๐ค`$equivariant sheaves descend naturally to the quotient $`_๐ค`$. To check that the cohomology of complexes $`(hR,๐_{Q_๐ค}|_\gamma )`$ and $`(hR,๐_{Q_๐ค}|_{\gamma ^\phi })`$ are canonically isomorphic it is sufficient to consider infinitesimal gauge transformation $`\phi =\epsilon \alpha `$, $`\epsilon ^2=0`$. Then the map $`Id+\epsilon m(\alpha )`$ gives the needed chain isomorphism. โ
Same statement holds true for arbitrary $`L_{\mathrm{}}`$modules over $`L_{\mathrm{}}`$algebra (in this case the vector field $`Q`$ and the $`Q`$connection $`๐_Q`$ will be given by formal power series and as usual one should express everything in terms of co-connections on comodules etc.).
Below we will specify this description in the case of the moduli space $``$ and the sheaf corresponding to the natural extension of the Gauss-Manin local system.
### 3.1. Local system on $`^{cs}`$.
Let us give first the description of the Gauss-Manin local system $`(^{cs},)`$ over the moduli space $`^{cs}`$. The sheaf $`^{cs}`$ is described by $`๐ค_{cs}`$module $`(\mathrm{\Omega }^,,\overline{}+)`$ with $`๐ค_{cs}`$action $`\gamma ๐ค_{cs}`$ $`_\gamma `$, $`_\gamma :=[,i_\gamma ]`$. This description proceeds as follows. Let $`\gamma MC_{cs}(R)`$ be a solution to eq.(2.1) and $`\varphi _\gamma :(SpecR,0)(^{cs},[X])`$ be the corresponding morphism representing a deformation of complex structures over $`R`$. The sheaf $`^{cs}`$ can be described via its inverse images $`\varphi _\gamma ^{}^{cs}`$. The sheaf $`\varphi _\gamma ^{}^{cs}\text{Spec}R`$ is given by $`R`$module defined as the cohomology of the complex
(3.2)
$$(\mathrm{\Omega }_R^,,d_\gamma ):=(\mathrm{\Omega }^,R,\overline{}++[i_\gamma ,])$$
here $`i_\gamma `$ denotes the operation of contraction with a vector field extended via wedge multiplication by $`(0,1)`$form. One can show easily that
$$\varphi _\gamma ^{}^{cs}=\underset{w}{}R^wf_{}(\mathrm{\Omega }_{\stackrel{~}{X}/\text{Spec}R}^{},d_{\stackrel{~}{X}/\text{Spec}R})$$
where $`f:\stackrel{~}{X}\text{Spec}R`$ is the flat family of complex structures on $`X`$ corresponding to $`\gamma `$. The sheaf $`^{cs}`$ is equipped with flat connection
(3.3)
$$:^{cs}^{cs}\mathrm{\Omega }_{^{cs}}^1,^2=0$$
Given a solution $`\gamma MC_{cs}(R)`$ the covariant derivative along a vector field $`vDer(R)`$ acting on element $`\phi `$ of $`\varphi _\gamma ^{}^{cs}`$ is written as
$$_v\phi =\frac{\phi }{v}+i_{\frac{\gamma }{v}}(\phi )$$
one can check that this is indeed the Gauss-Manin flat connection arising from the identification of de Rham cohomology with Betti cohomology.
### 3.2. Local system on $``$.
Notice that the $`๐ค_{cs}`$action on $`(\mathrm{\Omega }^,,\overline{}+)`$ extends naturally to the $`๐ค_A_{\mathrm{}}`$action $`\gamma _\gamma `$, $`_\gamma :=[,i_\gamma ]`$ where $`i_\gamma `$ denotes the operator of (holomorphic) contraction $`\mathrm{\Lambda }^{}T\mathrm{\Omega }^{}\mathrm{\Omega }^{}`$extended by wedge multiplication by $`(0,q)`$form. The description of the sheaf $``$ over $``$ follows the same scheme as above. Let us notice that the operator $`d_\gamma =\overline{}++[i_\gamma ,]`$ has the property $`d_\gamma ^2=0`$ for arbitrary element $`\gamma MC_A_{\mathrm{}}(R)`$ describing an $`A_{\mathrm{}}`$deformation $`\varphi _\gamma :\text{Spec}R`$. Therefore one can associate to such element $`\gamma `$ an $`R`$module $`{}_{}{}^{\prime \prime }\varphi _{\gamma }^{}^{\prime \prime }`$ equal to the cohomology of the complex $`(\mathrm{\Omega }_R^,,d_\gamma )`$. Also given two equivalent solutions $`\gamma _1,\gamma _2`$ the cohomology of the complexes $`(\mathrm{\Omega }_R^,,d_{\gamma _1}),(\mathrm{\Omega }_R^,,d_{\gamma _2})`$ are canonically isomorphic. It follows that the formula (3.2) with $`\gamma ๐ค_A_{\mathrm{}}`$ gives a sheaf $``$ .
To explain this in more details one has to recall first a definition of a notion of sheaf over formal moduli space described by functor (2.4). We will give a sketch of how one should proceed and leave the details to the reader. First of all let us consider functor $`\stackrel{~}{Def}_๐ค^{}`$ which associates to graded $``$Artin algebra $`R`$ the groupoid(=category where all morphisms are isomorphisms) whose objects are in one-to-one correspondence with elements $`\gamma MC(R)`$ and where morphisms $`\gamma _1\gamma _2`$ are in one-to-one correspondence with elements $`\stackrel{~}{\alpha }\mathrm{exp}\left((๐ค๐_R)^0\right)`$ transforming $`\gamma _1`$ to $`\gamma _2`$. Notice that the functor $`Def_๐ค^{}`$ can be obtained from $`\stackrel{~}{Def}_๐ค^{}`$ by taking $`\pi _0`$=the sets of connected components of objects. By definition a sheaf over moduli space described by $`Def_๐ค^{}`$ is a functor which associates to a graded $``$Artin algebra $`R`$ a representation of the grouppoid $`\stackrel{~}{Def}_๐ค^{}(R)`$ in the category of $`R`$modules. In our case it is the representation given by correspondence $`\gamma \varphi _\gamma ^{}`$.
The sheaf $``$ is equipped with flat connection defined by the same formula (3.3). Again formally this should be understand in terms of functors on Artin algebras similar to above.
In the sequel we will need a slight generalization of this situation involving formal parameter $`\mathrm{}`$. The fiber of the sheaf $`^{\mathrm{}}`$ over $`\varphi _\gamma `$ is given by cohomologies of complex
(3.4)
$$(\mathrm{\Omega }^,[[\mathrm{}^1,\mathrm{}]],\overline{}+[,i_\gamma ]+\mathrm{})$$
The covariant derivative along vector field $`vDer(R)`$ is given by the formula
(3.5)
$$_v=\frac{}{v}+\frac{1}{\mathrm{}}i_{\frac{\gamma }{v}}$$
### 3.3. $`๐`$modules over moduli spaces.
As a side remark let us mention that more generally, given a moduli space $``$ described by a differential graded Lie algebra $`๐ค`$, a sheaf $`H`$ equipped with flat connection ($`๐`$module) over $``$ can be described using the technique of deformation theory as follows. Let us denote via $`[\xi ]`$ the differential graded *commutative* algebra with generator $`\xi ,\xi ^2=0`$, $`\text{deg}\xi =1`$ and the differential $`d=\frac{}{\xi }`$. Notice that the space $`๐ค[\xi ]`$ is naturally a differential graded Lie algebra.
###### Proposition 3.2.
Given a $`๐ค`$module $`h`$ such that $`๐ค`$-module structure on $`h`$ is lifted to structure of a module over the differential graded Lie algebra $`๐ค[\xi ]`$ one has natural flat connection $``$ on the sheaf $`H`$ over $`_๐ค`$associated with $`h`$ according to prop.3.1. The covariant derivative along $`vDer(R)`$ acting on cohomology of $`๐_Q|_\gamma ,`$ $`\gamma MC_๐ค(R)`$ is given by $`_\nu =\frac{}{v}+i_{\frac{\gamma }{v}}`$where $`i_\alpha `$ denotes the action of the element $`\xi \alpha \epsilon ๐ค`$.
###### Proof.
It is straightforward to check that $`[,๐_Q]=0`$, $`[,]=0`$.
In our case $`๐ค_A_{\mathrm{}}`$module is the complex (3.4) with $`๐ค_A_{\mathrm{}}`$action $`\gamma ๐ค_A_{\mathrm{}}`$ $`_\gamma `$, $`_\gamma :=[,i_\gamma ]`$. The operators $`\frac{1}{\mathrm{}}i_\gamma `$ for $`\epsilon \gamma \epsilon ๐ค_A_{\mathrm{}}`$ define the lifting of this action to $`๐ค_A_{\mathrm{}}[\xi ]`$action. โ
###### Remark 3.3.
Let us mention that the same result holds true for $`L_{\mathrm{}}`$module structure over an $`L_{\mathrm{}}`$algebra $`๐ข`$ which can be lifted to $`L_{\mathrm{}}`$module structure over $`๐ข[\xi ]`$. The proof is the same with the only modification that $`[,]=2[๐_Q,S]`$ for some operator $`S`$. We see that $`L_{\mathrm{}}`$module over $`๐ข[\xi ]`$ describe higher homotopy generalization of the notion of sheaf with flat connection. There is in fact an infinite sequence of operators $`\widehat{}^0=๐_Q,`$ $`\widehat{}^1=`$, $`\widehat{}^2=S`$, $`\widehat{}^3`$$`\mathrm{}`$ satisfying higher homotopy identities. The existence of such higher homotopies allows one to define canonically a parallel transport operator acting on the complex $`(h,D_Q)`$ for the connection which is flat only up to homotopy. The details will appear in \[B2\].
## 4. Semi-infinite Hodge structure
In this section we define an analog of Hodge filtration for $`A_{\mathrm{}}`$deformations $`\varphi `$.
### 4.1. Grassmanian
For a $`_2`$graded $``$vector space $`H=(H^{\text{even}},H^{\text{odd}})`$, denote via $`H[[\mathrm{}^1,\mathrm{}]]`$ the space of formal series
$$\underset{i=\mathrm{}}{\overset{+\mathrm{}}{}}v_i\mathrm{}^i+\underset{j=\mathrm{}}{\overset{+\mathrm{}}{}}u_j\mathrm{}^{j+\frac{1}{2}},v_iH^{\text{even}},u_jH^{\text{odd}}$$
Denote also
$$H[[\mathrm{}]]=\{\underset{r0}{}a_r\mathrm{}^r\},H[[\mathrm{}^1]]=\{\underset{r<0}{}a_r\mathrm{}^r\}H[[\mathrm{}^1,\mathrm{}]]$$
and let $`H((\mathrm{}))`$ be the analogous ring of Laurent power series and $`p_+`$ be the linear projection $`H[[\mathrm{}^1,\mathrm{}]]H[[\mathrm{}]]`$ along $`H[[\mathrm{}^1]]`$.
Let $`\text{dim}_{}H<\mathrm{}`$. Let us introduce grassmanian $`Gr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])`$ of semi-infinite subspaces in $`H[[\mathrm{}^1,\mathrm{}]]`$. Firstly it contains grassmanian $`Gr_\frac{\mathrm{}}{2}\left(H((\mathrm{}))\right)`$ which consists of semi-infinite subspaces in $`H((\mathrm{}))`$. In fact, as we work in the neighborhood of moduli space of complex structures, we will be dealing only with the points of $`Gr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])`$ which lie in the formal neighborhood<sup>3</sup><sup>3</sup>3In general, for deformations which are not of finite order the subspaces which we study do not belong to $`Gr_\frac{\mathrm{}}{2}\left(H((\mathrm{}))\right)`$. One can repeat the same story in the analytic setting i.e. on the level of convergent solutions to Maurer-Cartan equation. In this case some type of Segal-Wilson grassmanian should be used. Then subspaces associated analogously with such convergent non-commutative deformations (see below) will not in general lie in $`H((\mathrm{}))`$ as well. of $`Gr_\frac{\mathrm{}}{2}\left(H((\mathrm{}))\right)`$. Under points of $`Gr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])`$ over (pro-)Artin algebra $`R`$ we will always mean in this paper points of such neighborhood. More precisely, a point $`LGr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])(R)`$ over an Artin algebra $`R`$ (i.e. a โfamily of subspacesโ represented by morphism $`SpecRGr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])`$) is an $`R[[\mathrm{}]]`$submodule $`LH((\mathrm{}))R`$ such that
$$L_{[[\mathrm{}]]}((\mathrm{}))=H((\mathrm{}))R$$
A point $`LGr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])(R)`$ over a pro-Artin algebra $`R`$ is a submodule in $`H[[\mathrm{}^1,\mathrm{}]]\widehat{}R`$ given by projective limit of submodules $`(proj)lim_iL_RR/๐_R^iGr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])(R/๐_R^i)`$ defined before. Such submodules always satisfy conditions $`dim_R\mathrm{ker}(p_+|_L)<\mathrm{}`$, $`dim_R\text{coker}(p_+|_L)<\mathrm{}`$.
Let $`F`$ be a decreasing filtration on $`H`$, $`\text{dim}_{}H<\mathrm{}`$. A filtration in the category of $`_2`$graded vector spaces is given by a pair of filtrations $`F=(F^{even},F^{odd})`$ on $`H^{even}`$and $`H^{odd}`$ respectively. Assume in addition that $`F^{even}`$ has indexes taking values in $``$ and $`F^{odd}`$ has indexes taking values in $`\frac{1}{2}+`$. To shorten the notations $`F^r`$will denote below the subspace $`(F^{even})^r`$ if $`r`$ and the subspace $`(F^{odd})^r`$ if $`r\frac{1}{2}+`$. To such filtration one can associate a subspace $`L^FGr_\frac{\mathrm{}}{2}(H[[\mathrm{}^1,\mathrm{}]])`$:
$$L^F:=\text{linear\hspace{0.17em} span}_{r[\frac{1}{2}]}F^r\mathrm{}^r[[\mathrm{}]]$$
The correspondence $`FL^F`$ is injective. In particular one has the subspace which corresponds to Hodge filtration<sup>4</sup><sup>4</sup>4we use indexes which are shifted with respect to the standard notations on $`H^{}(X,)`$
(4.1)
$$F^r(\varphi ):=\underset{pq2r,pq2r}{}H^{p,q},H^{p,q}=\{[\phi ]H^{p+q}(X,)[pq]|\phi \mathrm{\Omega }^{p,q}(\varphi )\}$$
which is associated with a deformation of complex structure $`\varphi ^{cs}`$.
### 4.2. Semi-infinite Hodge structure.
We claim that there is a subspace $`L(\varphi )Gr_\frac{\mathrm{}}{2}(H^{}(X,)[[\mathrm{}^1,\mathrm{}]])(R)`$ associated canonically to any $`A_{\mathrm{}}`$deformation $`\varphi (R)`$. For $`\varphi ^{cs}(R)`$ it coincides with the subspace $`L^{F(\varphi )}`$corresponding to the Hodge filtration (4.1).
The subspace $`L(\varphi )`$ is defined as follows. Let $`\gamma MC(R)`$ describes an $`A_{\mathrm{}}`$deformation $`\varphi (R)`$ where $`R`$ is an Artin algebra or a projective limit of Artin algebras. To such deformation we associate an $`R`$module $`L(\varphi )`$ in $`H^{}(X,)[[\mathrm{}^1,\mathrm{}]]R`$ representing morphism $`\text{Spec}RGr_\frac{\mathrm{}}{2}(H^{}(X,)[[\mathrm{}^1,\mathrm{}]])`$. Let $`l_{\mathrm{}}`$ be rescaling operator:
$$l_{\mathrm{}}:\phi ^{p,q}\mathrm{}^{\frac{qp}{2}}\phi ^{p,q}$$
Denote $`\stackrel{~}{L}_\gamma \mathrm{\Omega }^{}(X,)[[\mathrm{}^1,\mathrm{}]]R`$ the subspace of elements of the form<sup>5</sup><sup>5</sup>5The action of $`\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )`$ on arbitrary element from $`\mathrm{\Omega }^{}(X,)\widehat{}k[[\mathrm{}]]`$ is well-defined even in the case when $`R`$ is a pro-Artin algebra since $`\mathrm{exp}(\frac{1}{\mathrm{}}\gamma )`$ belongs to subspace of elements of the form $`_{j0}\gamma _j\gamma _j๐_R^j\mathrm{}^j๐ค_A_{\mathrm{}}\widehat{}k[[\mathrm{}]]`$.
(4.2)
$$l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(\phi _0+\phi _1\mathrm{}+\phi _2\mathrm{}^2\mathrm{})\text{where}\phi _i\mathrm{\Omega }^{}(X,)R$$
###### Proposition 4.1.
The subspace $`\stackrel{~}{L}_\gamma `$ is a sub-complex of the complex
(4.3)
$$(\mathrm{\Omega }^{}(X,)[[\mathrm{}^1,\mathrm{}]]R,\mathrm{}^{\frac{1}{2}}(\overline{}+))$$
###### Proof.
(4.4) $`l_{\mathrm{}}^1\mathrm{}^{\frac{1}{2}}(\overline{}+)l_{\mathrm{}}`$ $`=`$ $`(\overline{}+\mathrm{})`$
$`\mathrm{exp}({\displaystyle \frac{1}{\mathrm{}}}i_\gamma )(\overline{}+\mathrm{})\mathrm{exp}({\displaystyle \frac{1}{\mathrm{}}}i_\gamma )`$ $`=`$ $`\overline{}+\mathrm{}+{\displaystyle \frac{1}{\mathrm{}}}(i_{\overline{}\gamma +\frac{1}{2}[\gamma ,\gamma ]})+[,i_\gamma ]=`$
$`=`$ $`\overline{}+[,i_\gamma ]+\mathrm{}`$
Notice that the cohomology groups of the complex (4.3) are isomorphic to $`H^{}(X,)[[\mathrm{}^1,\mathrm{}]]R`$.
###### Definition.
The $`R`$submodule $`L(\varphi )H^{}(X,)[[\mathrm{}^1,\mathrm{}]]R`$ consists of cohomology classes represented by elements from $`\stackrel{~}{L}_\gamma `$.
###### Theorem 4.2.
To an arbitrary $`A_{\mathrm{}}`$deformation $`\varphi (R)`$ there is canonically associated subspace $`L(\varphi )Gr_\frac{\mathrm{}}{2}(H^{}(X,)[[\mathrm{}^1,\mathrm{}]])(R)`$ defined by formula (4.2). For $`\varphi ^{cs}(R)`$ this is the subspace $`L^{F(\varphi )}`$corresponding to the Hodge filtration (4.1).
###### Proof.
We need to check first that $`L(\varphi )=L(\varphi ^{})`$ for $`\gamma \gamma ^{}`$. It is enough to prove this for $`\gamma ^{}=\gamma +\epsilon (\overline{}\alpha +[\gamma ,\alpha ]),\epsilon ^2=0`$. Applying the standard commutation rules
$$[_{\gamma _1},i_{\gamma _2}]=i_{[\gamma _1,\gamma _2]},[i_{\gamma _1},i_{\gamma _2}]=0$$
one gets:
(4.6)
$$\mathrm{exp}(\frac{1}{\mathrm{}}i_{\gamma +\epsilon \overline{}\alpha +[\gamma ,\alpha ]})(\text{Id}+\epsilon [,i_\alpha ])=\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(\text{Id}+\epsilon [\overline{}+[,i_\gamma ]+\mathrm{},\frac{1}{\mathrm{}}i_\alpha ])$$
Assume that $`\phi \text{ker}(\overline{}+)\stackrel{~}{L}_\gamma `$, or equivalently
$$\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )l_{\mathrm{}}^1\phi \text{ker}(\overline{}+[,i_\gamma ]+\mathrm{})\mathrm{\Omega }^{}(X,)\widehat{}[[\mathrm{}]]$$
Applying the operator (4.6) to $`\stackrel{~}{\phi }=\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )l_{\mathrm{}}^1\phi `$ we get
$$\phi \pm \epsilon l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(\overline{}+[,i_\gamma ]+\mathrm{})\frac{1}{\mathrm{}}i_\alpha \stackrel{~}{\phi }\stackrel{~}{L}_{\gamma +\epsilon (\overline{}\alpha +[\gamma ,\alpha ])}$$
on the other hand
$$l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(\overline{}+[,i_\gamma ]+\mathrm{})=(\overline{}+)\mathrm{}^{\frac{1}{2}}l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )$$
Let $`\gamma \mathrm{\Omega }^{0,1}(X,T)๐_{R^0}`$ be a solution to Maurer-Cartan equation describing a deformation of complex structure $`\varphi :\text{Spec}R^{cs}`$. Let $`\{z^i\}`$ be a set of complex coordinates with respect to the fixed initial complex structure $`J_0`$. The differentials of the deformed complex coordinates can be written as $`(dz^i)^{\text{new}}=dz^i+_{\overline{j}}\gamma _{\overline{j}}^id\overline{z}^{\overline{j}}`$, where $`\gamma =_{i,\overline{j}}\gamma _{\overline{j}}^id\overline{z}^{\overline{j}}\frac{}{z^i}`$. Therefore $`F^r(\varphi )=\mathrm{exp}i_\gamma _{}F^r(J_0)`$. Notice also that for $`\gamma \mathrm{\Omega }^{0,1}(X,T)๐_{R^0}`$ one has $`l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )=\mathrm{exp}(i_\gamma )l_{\mathrm{}}`$. It follows that $`L(\varphi )=L^{F(\varphi )}`$ for $`\varphi ^{cs}(R)`$.
It is easy to see that there exist elements $`\alpha _j^k\mathrm{\Omega }^{}(X)`$, $`k=1,\mathrm{},\text{dim}H^{}(X,)`$, $`j=0,\mathrm{},m(k)`$, such that $`\overline{}\alpha _j^k=\alpha _{j1}^k`$, $`\overline{}\alpha _0^k=0`$, $`\alpha _{m(k)}^k=0`$ and such that $`\{_{j=0}^{m(k)}\alpha _j^k\}_{k=1,\mathrm{},\text{dim}H^{}(X,)}`$ give a basis in $`H^{}(X,)`$. Then given a solution to Maurer-Cartan equation $`\gamma MC_A_{\mathrm{}}(R)`$ over an Artin algebra $`R`$ describing $`\varphi _\gamma (R)`$ one has
$$\mathrm{}^N\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )\underset{j=0}{\overset{m(k)}{}}\alpha _j^k\mathrm{}^j\mathrm{\Omega }^{}(X)[[\mathrm{}]]$$
for sufficiently large $`N`$. Therefore $`l_{\mathrm{}}_{j=0}^{m(k)}\alpha _j^k\mathrm{}^j\mathrm{}^N\stackrel{~}{L}_\gamma `$ and $`L(\varphi _\gamma )_{[[\mathrm{}]]}((\mathrm{}))=H^{}(X,)((\mathrm{}))R`$
###### Remark 4.3.
The operator $`\frac{1}{\mathrm{}}i_\gamma `$ is an operator of the flat connection on the sheaf of cohomologies of complexes (3.4) described in the previous section. The operator $`l_{\mathrm{}}`$ identifies cohomology of complexes (3.4) and (4.3). We see that $`L(\varphi )`$ is the subspace obtained as a result of parallel transport to the base point $`[X]`$ of the subspace $`\{[\phi _0+\phi _1\mathrm{}+\phi _2\mathrm{}^2\mathrm{}]\}`$ living over $`\varphi `$ and identification provided by $`l_{\mathrm{}}`$.
###### Remark 4.4.
If $`X`$ is a compact Kahler manifold then $`\overline{}+[,i_\gamma ]`$ and $``$ satisfy $`\overline{}`$lemma. It follows that $`L(\varphi )`$ is a free $`R`$module for any $`\varphi _{[\gamma ]}(R)`$.
###### Remark 4.5.
As we already mentioned the moduli space $``$ can be viewed as a moduli space parameterizing $`A_{\mathrm{}}`$deformations of $`D^bCoh(X)`$. The local system over $``$ is the local system of periodic cyclic homology of the $`A_{\mathrm{}}`$categories and the semi-infinite subspace coincides with negative cyclic homology. Operator $`l_{\mathrm{}}`$ comes from natural grading on periodic cyclic homology of $`D^bCoh(X)`$.
## 5. Griffiths transversality
In this section we study variations of semi-infinite subspaces introduced in the previous section. Recall that the family of filtrations $`F^r(\varphi )`$ for deformations of complex structure $`\varphi ^{cs}`$ satisfies Griffiths transversality condition with respect to the Gauss-Manin connection:
$$F^rF^{r1}$$
###### Theorem 5.1.
The covariant derivative of $`L(\varphi )Gr_\frac{\mathrm{}}{2}(H^{}(X,)[[\mathrm{}^1,\mathrm{}]])(R),\varphi (R)`$ with respect to the Gauss-Manin connection satisfies
(5.1)
$$L\mathrm{}^1L$$
###### Proof.
Let $`\gamma MC(R)`$ describes an $`A_{\mathrm{}}`$deformation $`\varphi _\gamma (R)`$. Let
$$[\phi ]=[l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(\phi _0+\phi _1\mathrm{}+\mathrm{})]L(\varphi _\gamma )H^{}(X,)[[\mathrm{}^1,\mathrm{}]]R$$
describes a family of elements of the varying semi-infinite subspaces. Then for any vector field $`vDer(R)`$:
(5.2)
$$\begin{array}{c}_v[\phi ]=[l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(_v\phi _0+_v\phi _1\mathrm{}+\mathrm{})+\hfill \\ \hfill +\frac{1}{\mathrm{}}l_{\mathrm{}}\mathrm{exp}(\frac{1}{\mathrm{}}i_\gamma )(i_{_v\gamma }\phi _0+i_{_v\gamma }\phi _1\mathrm{}+\mathrm{})]\mathrm{}^1L(\varphi _\gamma )\end{array}$$
Notice that one has induced map (symbol of the Gauss-Manin connection)
(5.3)
$$T_\varphi L(\varphi )/\mathrm{}L(\varphi )\mathrm{}^1L(\varphi )/L(\varphi )$$
In the case of deformations of complex structure $`\varphi ^{cs}`$ this is the standard map
(5.4)
$$\left(\underset{p,q}{}H^q(X,\mathrm{\Lambda }^pT_X)[2]\right)\left(\underset{r}{}F^r/F^{r1}\right)\left(\underset{r}{}F^r/F^{r1}\right)$$
###### Remark 5.2.
Let us consider integrals
(5.5)
$$\phi ^i(t,\mathrm{})=_{\mathrm{\Delta }_i}\phi (t,\mathrm{})\text{for}\phi (t,\mathrm{})L(t),$$
where $`\{\mathrm{\Delta }_i\}`$ is a basis in $`H^{}(X,)`$ and $`t(R)`$ represents a class of $`A_{\mathrm{}}`$deformations of complex projective manifold $`X`$. Such integrals satisfy system of differential equations which are generalizations to the case of non-commutative deformations of Picard-Fuchs equations for usual periods. This can be seen as follows. Assume that we are given elements $`\phi _\alpha (0)F^{r_\alpha }(0),`$$`\phi _\alpha (0)F^{r_\alpha 1}(0)`$ which form a basis in $`H^{}(X,)`$. Let $`W_r`$ be an increasing $`[\frac{1}{2}]`$graded filtration opposite to filtration $`F^r(0)`$: that is, for any $`r`$, $`_{i2r}H^i(X,)=F^rW_r`$. Then $`L_W:=\text{linear\hspace{0.17em} span}_{r[\frac{1}{2}]}W_r\mathrm{}^r[[\mathrm{}^1]]`$ is a subspace transversal to $`L^{F(0)}`$. Let us consider elements $`\stackrel{~}{\phi }_\alpha (t,\mathrm{})=L(t)(\phi _\alpha (0)\mathrm{}^{r_\alpha }+L_W)`$. Notice that for $`t^{cs}`$, $`\stackrel{~}{\phi }_\alpha (t,\mathrm{})=\mathrm{}^{r_\alpha }\phi _\alpha (t)`$, where $`\phi _\alpha (t)=F^{r_\alpha }(t)(\phi _\alpha (0)+W_{r_\alpha })`$. Then one can show (see proof of prop.6.5 below for analogous arguments) that the periods $`\stackrel{~}{\phi }_\alpha ^i(t,\mathrm{})`$ satisfy the following system of equations: for any vector field $`vDer(R)`$ one has
$$\frac{\stackrel{~}{\phi }_\alpha ^i}{v}=\mathrm{}^1\underset{\beta }{}\mathrm{\Gamma }_v^\beta (t)\stackrel{~}{\phi }_\beta ^i$$
where $`\mathrm{}^1\mathrm{\Gamma }_v(t)`$ is the $`1`$form representing Gauss-Manin connection (3.5).
###### Remark 5.3.
One can generalize easily the notion of real polarized variation of Hodge structure to the semi-infinite context. For example, the semi-infinite analog of the standard condition<sup>6</sup><sup>6</sup>6here for reader convinience we use the standard grading on components of Hodge filtration
$$p,kF^p\overline{F}^{kp}=H^k$$
is given by
$$L(\mathrm{})\overline{L}(\overline{\mathrm{}}|_{\overline{\mathrm{}}=\mathrm{}^1})=H^{}(X,)[[\mathrm{}^1,\mathrm{}]]$$
satisfied by subspaces $`L(t)`$, $`t`$. Similarly, one can define semi-infinite analog of polarization form.
## 6. Quantum periods and counting of rational Gromov-Witten invariants on Calabi-Yau manifolds.
In three dimensions mirror symmetry predictions express rational Gromov-Witten invariants of a Calabi-Yau threefold in terms of variations of Hodge structure associated with the deformations of complex structure on the mirror dual Calabi-Yau threefold (see \[COGP\]). We explain in this section that the โnon-commutativeโ variations of semi-infinite Hodge structures described above play the same role in higher dimensional mirror symmetry. In this section $`X`$ denotes compact Kahler manifold with $`c_1(T_X)Pic(X)`$ equals to zero.
### 6.1. Moduli spaces of $`A_{\mathrm{}}`$deformations of Calabi-Yau manifolds
In the case when $`c_1(T_X)Pic(X)`$ equals to zero the moduli space $``$ has especially nice properties. Let us denote
$$๐=\underset{p,q}{}^{dim_{}H^q(X,\mathrm{\Lambda }^pT)}[pq]$$
Denote also via $`[[t_{}]]`$ the algebra of formal power series on the graded vector space $``$. It is convenient to fix some choice of a set $`\{t^a\}`$ of linear coordinates on $``$.
###### Proposition 6.1.
(\[BK\], lemma 2.1) The deformation functor $`Def_{๐ค_A_{\mathrm{}}}^{}`$describing the $`A_{\mathrm{}}`$deformations of $`X`$ is isomorphic to the functor represented by the pro-Artin algebra $`[[t_{}]]`$. Equivalently, there exists mini-versal solution to the Maurer-Cartan equation
$$\widehat{\gamma }(t)=\underset{a}{}\widehat{\gamma }_at^a+\frac{1}{2!}\underset{a_1,a_2}{}\widehat{\gamma }_{a_1a_2}t^{a_1}t^{a_2}+\mathrm{}MC([[t_{}]])$$
and the cohomology classes $`[\widehat{\gamma }_a]`$ form a basis in the cohomology of the complex $`(๐ค_A_{\mathrm{}},\overline{})`$.
###### Remark 6.2.
The reader shoud be warned that such solution is by no means unique. A choice of class of gauge equivalence of such solutions is equivalent to choice of system of local coordinates: $`(,0)(,[X])`$
### 6.2. Period map on $`^{cs}`$.
The condition $`c_1(T_X)=0`$ implies that there exists โholomorphic volume formโ $`\mathrm{\Omega }\mathrm{\Gamma }(X,\mathrm{\Omega }^n)`$ , $`n=\text{dim}_{}X`$, which is non-vanishing at any point of $`X`$. Such form is defined up to multiplication by a non-zero constant. Assume now that we have a family $`\stackrel{~}{X}_t`$ of deformations of complex structure on $`\stackrel{~}{X}_0=X`$. If we choose a holomorphic volume form $`\mathrm{\Omega }_0`$ on $`\stackrel{~}{X}_0`$ and a hyperplane $`lH^n(X,)`$ which is transversal to the line $`\mathrm{\Omega }_0`$ then the above mentioned ambiguity in the choice of holomorphic volume element can be fixed using the condition
$$\mathrm{\Omega }_t^l\mathrm{\Omega }_0l$$
for all values of the parameter $`t`$ for which the hyperplane $`l`$ rests transversal to the line $`\mathrm{\Omega }_t`$. The correspondence $`[\stackrel{~}{X}_t][\mathrm{\Omega }_t^l]`$ defines the period map: $`^{cs}H^n(X,)`$.
### 6.3. Period map on $``$.
In order to define analogous map on the moduli space $``$ let us fix a choice of subspace in $`H^{}(X,)[[\mathrm{}^1,\mathrm{}]]`$ transversal to $`L^{F(0)}`$ where $`F(0)`$ is the initial Hodge filtration on $`H^{}(X,)`$. Such subspace can be associated naturally with an increasing $`[\frac{1}{2}]`$graded filtration $`W=(W_{even},W_{odd})`$ on $`H^{}(X,)`$ which is opposite to $`F(0)`$ in the following sense:
(6.1)
$$r:\underset{i2r}{}H^i(X,)=F^rW_r$$
In the applications to mirror symmetry $`W`$ will be the limiting weight filtration associated with a maximal unipotency cusp on moduli space of complex structures on $`X`$. The subspace $`L_WH^{}(X,)[[\mathrm{}^1,\mathrm{}]]`$ associated with the filtration $`W`$ is defined in a similar manner as above for the decreasing filtration $`F^r`$ (see section 4.1)
$$L_W:=\text{linear\hspace{0.17em} span}_{r[\frac{1}{2}]}W_r\mathrm{}^r[[\mathrm{}^1]]$$
where we assumed as above that $`W_r:=W_r^{even}`$ for $`r`$, and $`W_r:=W_r^{odd}`$ for $`r\frac{1}{2}+`$. It follows from the condition (6.1) that $`L_W`$ is transversal to $`L^{F(0)}`$:
(6.2)
$$H^{}(X,)[[\mathrm{}^1,\mathrm{}]]=L_WL^{F(0)}$$
Assume now that in addition to filtration $`W`$ satisfying (6.1) a holomorphic volume element $`\mathrm{\Omega }_0`$ for initial complex structure on $`X`$ is fixed. Let $`\gamma MC_A_{\mathrm{}}(R)`$ represents an $`A_{\mathrm{}}`$deformation $`\varphi :(R).`$ Recall that we denoted via $`L(\varphi )H^{}(X,)[[\mathrm{}^1,\mathrm{}]]R`$ an $`R`$submodule representing the family of semi-infinite subspaces.
###### Proposition 6.3.
The intersection $`L(\varphi )(\mathrm{\Omega }_0\mathrm{}^{\frac{n}{2}}+L_WR)`$ consists of a single element.
###### Proof.
It follows from the condition (6.1) that $`L_W`$ is transversal to $`L^{F(0)}`$. One can show that $`R`$module $`L(\varphi )`$ is free (it follows, for example, from \[B1\], prop. 4.2.1). Therefore $`H^{}(X,)[[\mathrm{}^1,\mathrm{}]]R=L(\varphi )(L_WR)`$. โ
We will denote the element defined by intersection $`L(\varphi _\gamma )(\mathrm{\Omega }_0\mathrm{}^{\frac{n}{2}}+L_WR)`$ via $`\mathrm{\Psi }^W(\varphi ,\mathrm{})`$. For a mini-versal solution $`\widehat{\gamma }MC([t_{}])`$ this element coincides with the series $`(\mathrm{\Phi }^W(t_{},\mathrm{})\mathrm{\Omega }_0+\mathrm{\Omega }_0)\mathrm{}^{\frac{n}{2}}`$ from \[B1\] (here $`\mathrm{\Phi }^W(t_{},\mathrm{})`$ is the series defined in eq. (4.28) after theorem 1 in \[B1\]). The corresponding morphism
$$\mathrm{\Psi }^W(\varphi _{\widehat{\gamma }},\mathrm{}):H^{}(X,)[[\mathrm{}^1,\mathrm{}]][n(\text{mod})2]$$
does not depend on the choice of the mini-versal solution $`\widehat{\gamma }`$.
###### Proposition 6.4.
(\[B1\] propositions 4.2.4 and 4.2.6) The map $`\mathrm{\Psi }^W(\varphi _{\widehat{\gamma }},\mathrm{})`$ has the following properties:
* the restriction $`\mathrm{\Pi }^W:=`$ $`\mathrm{\Psi }^W(\mathrm{})|_{\mathrm{}=1}`$is well-defined and gives a local isomorphism of formal super-manifolds (local Torelli theorem):
$$\mathrm{\Pi }^W:(,[X])(H^{}(X,)[n(\text{mod})2],[\mathrm{\Omega }_0])$$
* the restriction $`\mathrm{\Pi }^W|_{^{cs}}`$ coincides with the classical period map $`[X_t][\mathrm{\Omega }_{[X_t]}^{W_{\frac{n}{2}}}]H^n(X,)`$ defined with help of the hyperplane $`W_{\frac{n}{2}}H^n(X,)`$.
Let us consider โvalue at $`\mathrm{}=\mathrm{}`$โ of $`\mathrm{\Psi }^W`$ which is a morphism $`\mathrm{\Psi }^W(\mathrm{}):GrW`$ into associated quotient of the filtration $`W`$ defined as the composition of $`\mathrm{\Psi }^W`$ with affine map $`\mathrm{\Omega }_0\mathrm{}^{\frac{n}{2}}+L_WL_W/\mathrm{}^1L_W=_rW_r/W_{r1}`$ sending $`v=\mathrm{\Omega }_0\mathrm{}^{\frac{n}{2}}+v_r\mathrm{}^r+v_{r+1}\mathrm{}^{r+1}+\mathrm{}\mathrm{\Omega }_0\mathrm{}^{\frac{n}{2}}+W_r\mathrm{}^r[[\mathrm{}^1]]`$ to $`[v_r]W_r/W_{r1}`$. This morphism is a local isomorphism as well (\[B1\], eq.(4.25)) and defines a set of local coordinates on $``$ which we denote by $`t_W`$.
### 6.4. Mirror symmetry in higher dimensions.
The map
$$\left(\underset{p,q}{}H^q(X,\mathrm{\Lambda }^pT_X)[pq(\text{mod})2]\right)[\mathrm{\Omega }_0]\left(\underset{r}{}F^r/F^{r1}\right)[n(\text{mod})2]$$
given by restriction of the symbol of Gauss-Manin connection (5.4) is an isomorphism. Therefore the set $`\{[\frac{\mathrm{\Psi }^W(t_{},\mathrm{})}{t_{}^a}]\text{mod}L([\widehat{\gamma }])\mathrm{}^1L([\widehat{\gamma }])/L([\widehat{\gamma }])\}`$ is a set of free generators of $`\left(\mathrm{}^1L([\widehat{\gamma }])\right)/L([\widehat{\gamma }])`$. This implies that the quantum periods given by components of the map $`\mathrm{\Psi }^W`$ satisfy a system of differential equations (non-commutative extension of a version of Picard-Fuchs equations):
###### Proposition 6.5.
The series $`\mathrm{\Psi }^W(t_W,\mathrm{})`$ satisfy:
$$\frac{^2\mathrm{\Psi }^W}{t_W^at_W^b}=\mathrm{}^1\underset{c}{}A_{ab}^c(t_W)\frac{\mathrm{\Psi }^W}{t_W^c}$$
for some $`A_{ab}^c(t_W)[[t_W]]`$ where $`\{t_W\}`$ is a set of linear coordinates on $`GrW`$ induced by the map $`\mathrm{\Psi }^W(\mathrm{})`$.
###### Proof.
It follows from theorem 5.1 that
$$\frac{^2\mathrm{\Psi }^W}{t_{}^at_{}^b}\mathrm{}^2L([\widehat{\gamma }])$$
Therefore there exist $`\left(A^{(1)}\right)_{ab}^c(t_{}),\left(A^{(0)}\right)_{ab}^c(t_{})[[t_{}]]`$ such that
$$\frac{^2\mathrm{\Psi }^W}{t_{}^at_{}^b}\mathrm{}^1\underset{c}{}\left(A^{(1)}\right)_{ab}^c(t_{})\frac{\mathrm{\Psi }^W}{t_{}^c}\underset{c}{}\left(A^{(0)}\right)_{ab}^c(t_{})\frac{\mathrm{\Psi }^W}{t_{}^c}L([\widehat{\gamma }])$$
On the other hand,
$$\frac{^2\mathrm{\Psi }^W}{t_{}^at_{}^b},\frac{\mathrm{\Psi }^W}{t_{}^a}L_W\widehat{}[[t_{}]]$$
Therefore
$$\frac{^2\mathrm{\Psi }^W}{t_{}^at_{}^b}=\mathrm{}^1\underset{c}{}\left(A^{(1)}\right)_{ab}^c(t_{})\frac{\mathrm{\Psi }^W}{t_{}^c}+\underset{c}{}\left(A^{(0)}\right)_{ab}^c(t_{})\frac{\mathrm{\Psi }^W}{t_{}^c}$$
The coordinates $`\{t_W\}`$ were chosen so that $`\mathrm{\Psi }^W\text{mod}\mathrm{}^1L_W:L_W/\mathrm{}^1L_W`$ is linear in $`\{t_W\}`$. Therefore in these coordinates $`\frac{^2\mathrm{\Psi }^W}{t_W^at_W^b}=0\text{mod}\mathrm{}^1L_W`$ and $`\left(A^{(0)}\right)_{ab}^c=0`$. โ
###### Corollary 6.6.
One has for one-form $`A=_aA_{ab}^c(t_W)dt_W^a`$ :
(6.3)
$$dA=0,[A,A]=0$$
The map $`\mathrm{\Psi }^W`$ induces from the Poincare pairing a bilinear form on the tangent sheaf of $``$ ( \[B1\] formula (5.59)). If $`W`$ is isotropic with respect to the Poincare pairing:
$$(\alpha ,\beta )=0,\text{for}\alpha W_r,\beta W_{r+1}$$
then the form induced on $`T`$ via $`\mathrm{\Psi }^W`$is written as $`\frac{}{t_W^a},\frac{}{t_W^b}=\eta _{ab}\mathrm{}^{n2}`$, where $`\eta _{ab}`$ is graded symmetric and non-degenerate. The series $`A_{ab}^c(t_W)`$ give structure constants of commutative associative multiplication on the fibers of $`T`$ ($`[A,A]=0`$). They define together with $`\eta _{ab}`$ a quasi-homogeneous solution to WDVV-equation (\[B1\] theorem 5.6) on (formal) neighborhood of zero in $`GrW`$ identified with $``$.
###### Remark 6.7.
Similarly one can construct solutions to WDVV-equation starting from a of kind of abstract semi-infinite variation of Hodge structure of Calabi-Yau type. More precisely it is the data $`(๐ข,\mathrm{\Omega }[\mathrm{}])`$ where $`๐ข`$ is an $`L_{\mathrm{}}`$algebra and $`\mathrm{\Omega }[\mathrm{}]`$ is $`[\mathrm{}]`$linear $`L_{\mathrm{}}`$module over $`๐ข^{\mathrm{}}[\xi ]`$, here $`^{\mathrm{}}[\xi ]`$ is the differential graded commutative algebra over $`[\mathrm{}]`$ with generator $`\xi ,\xi ^2=0`$, $`\text{deg}\xi =1`$ and the differential $`d=\mathrm{}\frac{}{\xi }`$. This data define local system $`^{\mathrm{}}`$ over the moduli space associated with $`๐ข`$ and the family of semi-infinite subspaces $`L(t)^{\mathrm{}}(t)`$ $`t`$, which satisfy analog of Griffiths transversality condition (5.1). We assume that moduli space $``$ has a smooth subspace $`\stackrel{~}{}`$ such that $`L(t)|_\stackrel{~}{}`$ is described by a flat $`๐ช_\stackrel{~}{}`$ module and that Calabi-Yau type condition holds on cohomology level: there exists $`[\mathrm{\Omega }_0]L(0)/\mathrm{}L(0)`$, such that the restriction of symbol of Gauss-Manin connection (5.3) to $`[\mathrm{\Omega }_0]`$ gives an isomorphism $`T_0\stackrel{~}{}\mathrm{}^1L(0)/L(0)`$. Then to any subspace $`S_0^{\mathrm{}}`$, $`\mathrm{}^1SS`$ transversal to $`L(0)`$ one can associate a solution $`A_{ab}^c(t)`$ to equations (6.3). Assume in addition that the module $`\mathrm{\Omega }[\mathrm{}]`$ is equipped with non-degenerate invariant pairing $`G(\mathrm{}a,b)=G(a,\mathrm{}b)=\mathrm{}G(a,b)`$ which induces pairing on $`L(0)`$ with values in $`\mathrm{}^D[[\mathrm{}]]`$ for some $`D`$. Then for isotropic $`S`$ ($`G|_\mathrm{}S\mathrm{}^D[[\mathrm{}^1]]`$) a flat metrics is induced on $`\stackrel{~}{}`$ compatible with multiplication defined by $`A_{ab}^c(t)`$, i.e. one gets a solution to WDVV equation.
###### Remark 6.8.
Using the previous remark one can repeat the above construction of solutions to WDVV-equation starting from differential graded Lie algebra of Hochcshild cochains describing $`A_{\mathrm{}}`$deformations of $`A_{\mathrm{}}`$category $`D^bCoh(X)`$ and the dg-module describing variations of semi-infinite Hodge structure associated with local system of periodic cyclic homology. Consequently, the above formal power series $`A_{ab}^c(t_W)`$ can be constructed entirely in terms of $`A_{\mathrm{}}`$category $`D^bCoh(X)`$.
Rational Gromov-Witten invariants of Calabi-Yau manifold $`Y`$ are encoded in the series $`C_{ab}^c(t_H,q)^Y`$ of structure constants of (big) quantum cohomology (see \[KM\]). In \[B1\] we made a conjecture which can be now reformulated using the fact that the power series $`\mathrm{\Pi }^W`$ introduced above coincides with analogous power series from \[B1\]. The conjecture states that rational Gromov-Witten invariants of $`Y`$ coincide with Taylor coefficients in the series
$$\underset{\alpha }{}\frac{^2\mathrm{\Pi }^{W,\alpha }}{t_W^at_W^b}\left((\mathrm{\Pi }^W)^1\right)_\alpha ^c=A_{ab}^c(t_W)^X$$
where $`\mathrm{\Pi }^{W,\alpha }(t_W)=_{\mathrm{\Delta }_\alpha }\mathrm{\Pi }^W`$ are the quantum periods considered as functions on the moduli space $``$ associated with the mirror dual family of Calabi-Yau manifolds $`X`$. Some care is needed here: technically we consider family of expansions of $`A_{ab}^c(t_W,q)^X`$ with varying base $`[X_q]^{cs}`$, $`[X_q]`$ is close to point of maximal unipotent monodromy on the boundary of moduli space of complex structures on $`X`$. This is in agreement with the fact that $`C_{ab}^c(t_H,q)^Y`$ are formal power series with coefficients in semi-group ring $`Q[B]`$ where $`B`$ is the semi-group of algebraic cycles on $`Y`$ modulo numerical equivalences. Interested reader is referred to \[B1\] ยง6 for details. Explicitly, the conjecture now states that the identity
(6.4)
$$C_{ab}^c(t_H,q)^Y=A_{ab}^c(t_W,q)^X$$
should hold, where $`W`$ is the limiting weight filtration on $`H^{}(X,)`$ associated with a maximal unipotency cusp on moduli space of complex structures on $`X`$ and appropriate choice of identification of affine spaces $`H^{}(Y,)=GrW`$ is assumed. We checked in \[B1\] this conjecture for projective complete intersections:
###### Theorem.
(\[B1\] theorem 6.2) The identity $`C_{ab}^c(t_H,q)^Y=A_{ab}^c(t_W,q)^X`$ holds for projective complete intersection Calabi-Yau varieties and their mirrors.
###### Remark 6.9.
One can show using prop. 6.4b that in dimensions three this is equivalent to the standard predictions from \[COGP\] proven in \[G\]. In \[G\] some partial higher dimensional generalization is also proven, which deals with ordinary periods and the subset of rational Gromov-Witten invariants corresponding to small quantum cohomology ring.
### 6.5. Quantum periods and mirror symmetry for Fano manifolds.
A somewhat similar type of integrals appears in a different setting, namely as periods associated with $`A_{\mathrm{}}`$deformations of a pair $`(X,f)`$ where $`X`$ is a quasi-projective variety and $`f:X๐ธ^1`$ is a morphism to affine line. Such objects appear as mirror partners to projective complete intersection Fano manifolds. We consider such a case in \[B3\]. In fact in this case the situation simplifies drastically. The Maurer-Cartan equation becomes empty, the corresponding moduli space can be identified with the moduli space of deformations of the function $`f`$ and the periods of the semi-infinite Hodge structures are given simply by oscillating integrals. We prove in loc.cit. an identity analogous to (6.4) which expresses total collection of rational Gromov-Witten invariants of $`^n`$ in terms of periods of semi-infinite Hodge structure associated with its mirror partner.
|
warning/0006/astro-ph0006169.html
|
ar5iv
|
text
|
# The LBDS Hercules sample of millijansky radio sources at 1.4 GHz: I. Multi-colour photometry
## 1 Introduction
Essentially all low-redshift, powerful ($`\mathrm{P}_{1.4\mathrm{GHz}}\mathrm{P}_{1.4\mathrm{GHz}}^{}=10^{25}`$ W Hz<sup>-1</sup> sr<sup>-1</sup>) radio galaxies are housed in normal giant elliptical galaxies. Thus radio selection has long been regarded as providing an efficient method of identifying high-redshift ellipticals. However, very radio-powerful high-redshift sources have proved virtually useless as probes of galaxy evolution due to the contaminating effects of the active galactic nucleus (AGN) at ultravioletโinfrared wavelengths (McCarthy et al. 1987; Tadhunter et al. 1992; Best, Longair & Rรถttgering 1996, 1997). Given the success of the Lyman-break technique for selecting high-redshift galaxies on the basis of their optical colours (Steidel et al. 1996, 1999), it might be presumed that radio galaxies are no longer of any importance for probing galaxy evolution. In fact, nothing could be further from the truth, since the identification of high-redshift galaxies at optical wavelengths is by necessity biased towards the bluest, most ultraviolet-active systems โ such methods cannot detect old or dust-reddened sources. What is required is an optically unbiased selection method, such as radio selection at low flux densities. Unlike their high-power counterparts, faint radio sources have only weak emission lines (Kron, Koo & Windhorst 1985) and show little evidence of an alignment effect (Dunlop & Peacock 1993).
There are a number of radio surveys with millijansky flux density limits at 1.4 GHz ($`S_{1.4}1`$ mJy). This limit is sufficiently faint that a large fraction of the sources potentially lie at high redshifts, and yet sufficiently bright that objects selected at this flux density are still predominantly radio galaxies rather than low-redshift starbursts (Hopkins et al. 2000). Some of the more recent surveys are the FIRST survey (Becker, White & Helfand 1995; White et al. 1997), the NVSS (Condon et al. 1998), the VLA survey of the ELAIS regions (Ciliegi et al. 1999), and the ATCA survey of the Marano Field (Gruppioni et al. 1997). In order to study the evolution of the hosts of these radio sources, it is essential to identify optical/infrared counterparts to a majority of the radio sources. This has not yet been achieved for any of these recent surveys. This paper presents essentially complete optical identifications for a subsample of the LeidenโBerkeley Deep Survey (LBDS; Windhorst, van Heerde & Katgert 1984a), together with infrared $`K`$-band observations of 85% of the optically identified sources. A companion paper (Waddington et al. 2000a; hereafter Paper II) will present the results of the spectroscopic observations and photometric redshift estimates of these sources.
The layout of the paper is as follows. In section 2 the radio observations and previous work using the LBDS are reviewed, and the resolution corrections for the sources are recalculated. In section 3 the optical identifications and photometry of sources in the Hercules sample are presented. The infrared observations are presented in section 4. Finally, section 5 is a discussion of the properties of the sample as a whole and of a number of the individual sources, and compares the $`K`$-band magnitude distribution to that of brighter radio surveys. Unless otherwise stated, a cosmology with $`H_0=50`$ km s<sup>-1</sup> Mpc<sup>-1</sup>, $`\mathrm{\Omega }_0=1`$ and $`\mathrm{\Lambda }=0`$ is assumed.
## 2 The LeidenโBerkeley Deep Survey
The LBDS began with a collection of multi-colour prime focus plates that had been acquired with the 4-m Mayall Telescope at Kitt Peak for the purpose of faint galaxy and quasar photometry \[Kron 1980, Koo & Kron 1982\]. Several high latitude fields were chosen in the selected areas SA28 (referred to as Lynx), SA57, SA68 and an area in Hercules. The fields were selected on purely optical criteria, such as high galactic latitude ($`b^{II}35\mathrm{ยฐ}`$) and minimum H i column density, resulting in a random area of sky from the radio perspective. Nine of these fields were then surveyed with the 3-km Westerbork Synthesis Radio Telescope (WSRT) at 21 cm (1.412 GHz) with a 12$`\stackrel{}{.}`$5 beam \[Windhorst et al. 1984a\]. These observations yielded noise limited maps with $`\sigma =0.12`$โ0.28 mJy in 12 hours, with absolute uncertainties in position of order 0$`\stackrel{}{.}`$4 and in flux of order 3โ5%. A total of 471 sources were found, out of which a complete sample of 306 sources was selected. This complete sample is defined by: (1) a peak signal-to-noise $`S_\mathrm{p}/N5\sigma `$, corresponding to $`S_\mathrm{p}0.6`$โ1.0 mJy; and (2) a limiting primary beam attenuation of $`7`$ dB, which corresponds to a radius of 0$`\stackrel{}{.}`$464, or an attenuation factor $`A(r)5`$. With this attenuation radius, the WSRT field then covers the same area as the Mayall 4-m plates. Images were also made at the resolution of the 1.5-km WSRT ($`\sigma =0.27`$ mJy; Windhorst et al. 1984a) and at 50 cm (0.609 GHz) with the 3-km WSRT ($`\sigma =0.48`$ mJy; Windhorst 1984). The radio data for the two Hercules fields are summarised in Table 1.
Following this selection of the radio sample, the photographic plates were searched for optical counterparts to the radio sources (Windhorst, Kron & Koo 1984b). Large astrographic plates were used to measure the positions of 30โ40 secondary standards based on the positions of AGK-3 astrometric standard stars. These secondary standards were then used to calibrate the Mayall prime focus plates for each field in the survey. The radio and optical position errors combine to give an error ellipse around the co-ordinates of the radio source within which one expects to find the optical identification of the source. For the complete sample, 171 of the 306 radio sources were identified in this way, to $`U^+`$, $`J^+23.5`$ mag and $`F^+`$, $`N^+21.5`$โ22.5 mag. The expected number of real identifications was 53% for the whole survey, while for the Hercules fields the identification percentage was somewhat higher at 65% (47 out of 72 sources). As will be discussed in Paper II, there is some evidence for large-scale structure in Hercules, which may explain the higher identification rate in these two fields.
Magnitudes for each of these objects were measured from digital scans of the plates \[Kron et al. 1985\]. Absolute calibrations were made for most of the fields using photoelectric observations of stars in the fields and transforming the derived $`UBV`$ zero-points to the photographic $`U^+J^+F^+N^+`$ system. For the two Hercules fields, however, no photoelectric zero-points were available. They had to be estimated by assuming that the sensitivity of the Hercules plates was the same as for plates (with the same emulsion and filter) taken of another field for which photometric calibrations were available. The resulting errors in the photometry (0.2โ0.3 mag) are thus larger for Hercules than the rest of the sample. Redshifts were measured for approximately 60 of the LBDS identifications, of which 16 were in Hercules \[Kron et al. 1985\].
One of the major limiting factors in using the LBDS for statistical studies of radio source evolution has been the incomplete optical identification of the sample and the limited number of redshifts available. To obtain complete identification and spectroscopic data on the whole 306-source survey was a major long-term project, due to the size of the survey area (several square degrees), so here attention is focused on a subsample of 72 sources. The two Hercules fields were chosen for this detailed study because they had the largest number of optical identifications already available. In addition, the highest redshift object yet known in the LBDS (53W002 at $`z=2.390`$) was found in these fields. Optical and infrared observations have been acquired over many years and are presented here for the first time in their entirety. First, some of the relevant highlights of previous work in the LBDS will be reviewed and the radio observations of Hercules will be summarised.
### 2.1 Previous work in the LBDS
Results from the photographic identifications were presented by Kron et al. (1985). About 20% of the identifications were quasars and there were two stars. The galaxies were separated into bright ($`F^+<18`$) and faint ($`18<F^+<21.5`$), red ($`J^+F^+>1.2`$โ2.8) and blue ($`J^+F^+<1.2`$โ2.8) classes. The red radio galaxies (both bright & faint classes) were identified as giant ellipticals with extended radio emission โ โclassicalโ powerful radio galaxies. The bright, blue galaxies were exclusively spirals with unresolved radio structure and a very narrow range in power $`\mathrm{log}_{10}P_{1.4}=22.1\pm 0.1`$. The faint, blue galaxies frequently had irregular optical morphologies (interactions, mergers or compact galaxies), and higher radio/optical flux ratios than normal spirals, suggesting they were a population distinct from the brighter blue sources.
Thuan et al. (1984) observed 48 sources in the LBDS in the near-infrared $`JHK`$-bands, to $`K17.5`$ mag. They discovered that the opticalโinfrared colours of the faint radio galaxies were due to their stellar populations, showing no correlation with their radio flux or power. Less than 10% of the red radio galaxies had evidence for non-thermal infrared emission. The colours of the red galaxies were consistent with those of distant ($`z0.2`$) non-evolving or passively-evolving luminous giant ellipticals, such as observed in the 3CR survey \[Lilly & Longair 1984\]. The optical/infrared colours of the faint blue radio galaxies were indicative of low-luminosity and/or star-forming galaxies.
High-resolution maps of many of the LBDS sources were obtained at 1.4 GHz with the Very Large Array (VLA), in order to investigate the radio morphologies of weak radio sources \[Oort et al. 1987, Oort 1988\]. The median angular size of the radio samples was found to decrease with decreasing flux density, to $`2`$โณ at 1โ10 mJy and $`1`$โณ at 0.4โ1.0 mJy. This was attributed to the combined effects of a change in the population and a decrease in the intrinsic size of radio sources at lower (radio) luminosity. Approximately 90% of the red galaxies were resolved by the 1$`\stackrel{}{.}`$4 beam, but half of the blue radio galaxies remained unresolved. There was evidence in the sample for a decrease in intrinsic size with increasing redshift โ a factor of 4 from $`z=0`$ to $`z=0.5`$. The morphological information and radio positions from this survey have been incorporated into Table 1. Radio observations have also been made of the LBDS at additional frequencies, for example: 4.86 GHz (6 cm) by Donnelly, Partridge & Windhorst (1987), 327 MHz (92 cm) by Oort, Steemers & Windhorst (1988). Deeper VLA observations were also made of the Lynx-2 field, reaching 5-$`\sigma `$ limiting fluxes of 0.2 mJy at 1.41 GHz \[Windhorst et al. 1985\] and $`20`$ $`\mu `$Jy at 8.44 GHz \[Windhorst et al. 1993\].
Several sources drawn from the LBDS Hercules sample have also been the subject of individual investigations. The most widely-studied of these is 53W002, which, at $`z=2.390`$, is the most distant source to date in the whole LBDS \[Windhorst et al. 1991\]. A total of 67 orbits with the Hubble Space Telescope (HST) have now been devoted to 53W002 and its companions. Windhorst, Keel & Pascarelle (1998) used these data to investigate the alignment effect and evolution of this relatively weak high-redshift radio galaxy. Narrow-band images of the field around the galaxy led Pascarelle et al. (1996a,b) to discover a group of emission-line objects at $`z2.4`$ around the radio source, which they suggested were subgalactic clumps that would eventually collapse into a few massive galaxies today.
Spectroscopic observations with the Keck Telescope of two of the reddest sources (53W069 & 53W091) have enabled detailed models of their spectra to be used to estimate their ages. 53W091 was found to have a likely age of 3.5 Gyr at $`z=1.552`$ \[Dunlop et al. 1996, Spinrad et al. 1997\], and a best-fitting age of 4.0 Gyr was determined for 53W069 at $`z=1.432`$ \[Dey 1997, Dunlop 1999\]. However, other authors have suggested that the age of 53W091 is nearer to 1.5โ2.0 Gyr \[Bruzual & Magris 1997, Yi et al. 2000\]. In addition to 53W002, HST observations of several other LBDS Hercules sources have been published: 53W044 & 53W046 \[Keel & Windhorst 1993, Windhorst et al. 1994\] and 53W069 & 53W091 \[Waddington et al. 2000b\].
### 2.2 Sample completeness and source weights
Before reviewing the radio properties of the LBDS Hercules sample, it is necessary to discuss the completeness of the sample and the associated weighting of radio sources. Given the selection criteria of the LBDS ($`S_\mathrm{p}/N5\sigma `$, $`A(r)5`$), the sample will be complete only for certain ranges in total flux density $`S_\nu `$, angular size $`\psi `$ and the component flux ratio $`f`$. Here, we summarize the discussion of the completeness, that is given in full in Windhorst et al. (1984a).
Due to the primary beam attenuation (i.e. the sensitivity of the telescope decreases with radial distance from the pointing centre), the observed signal-to-noise of a given source will depend on where it is on the map. For example, a source that lies just above the 5-$`\sigma `$ cut at the centre of the map would have fallen below the cut if it happened to lie at the edge, and it would then not be included in the sample. Assuming the primary beam is isotropic, then the correction for this incompleteness is a simple, geometrically-determined weight in the source counts and optical identification statistics. Every radio source is weighted inversely proportional to the area over which it could have been seen, before it dropped below the 5-$`\sigma `$ threshold.
This primary beam weight is
$$W_{\mathrm{pb}}=\left(\frac{0.464}{r_{\mathrm{lim}}}\right)^2$$
(1)
where $`r_{\mathrm{lim}}`$ is the distance from the beam centre (in degrees) at which the source would fall below the 5-$`\sigma `$ cut:
$$r_{\mathrm{lim}}=\frac{1}{61.191\times \nu }\mathrm{arccos}\left[\left(\frac{S_\mathrm{p}}{N}\right)\frac{A(r_{\mathrm{obs}})}{5}\right]^{\frac{1}{6}}.$$
(2)
Here, $`\nu `$ is the frequency in GHz, $`S_\mathrm{p}/N`$ is the observed peak signal-to-noise, and the attenuation ($`A`$) at the distance ($`r_{\mathrm{obs}}`$) from the beam centre where the source was observed is given by Windhorst et al. (1984a) as
$$A(r_{\mathrm{obs}})=\mathrm{cos}^6\left(61.191\times \nu \times r_{\mathrm{obs}}\right).$$
(3)
A second, more complicated, source of incompleteness is the resolution/population bias โ a resolved source of a given sky flux $`S_{^{\mathrm{TOT}}}^{_{\mathrm{SKY}}}`$ will drop below the 5-$`\sigma `$ cut-off more easily than a point source of the same $`S_{^{\mathrm{TOT}}}^{_{\mathrm{SKY}}}`$. Thus the completeness depends on the distribution of source angular sizes and component flux ratios. The determination of, and statistical corrections for, this incompleteness were found by processing Monte Carlo simulations of artificial data using the same algorithms as the real data. They are detailed in Windhorst et al. (1984a), to which one is referred. Higher resolution (arcsecond) VLA images have subsequently shown that the median angular size of radio sources is a function of flux density, decreasing monotonically from $`20`$โณ at 1 Jy to $`2`$โณ at 1 mJy. Windhorst, Mathis & Neuschaefer (1990) showed that this resulted in an overestimate of the original resolution correction, which was based on the lower resolution WSRT data, by a factor of approximately twenty-five percent. The revised source weights were computed following equation (14) of Windhorst et al. (1984a), incorporating the revised angular sizes, as follows:
$$W_{\mathrm{res}}=1+\frac{58\mathrm{}}{\mathrm{HPBW}_\alpha }\mathrm{exp}\left(\frac{2.71(S_{\mathrm{TOT}}/N)}{(S_\mathrm{p}/N)_{\mathrm{cut}\mathrm{off}}}\right).$$
(4)
Here, $`\mathrm{HPBW}_\alpha =12\mathrm{}`$ is the half-power beam-width of the observations, $`S_{^{\mathrm{TOT}}}/N`$ is the total observed signal-to-noise of the source and $`(S_\mathrm{p}/N)_{\mathrm{cut}\mathrm{off}}=5`$ is the selection criterion of the sample. The statistical errors in this correction were shown to be a few percent at the $`S_{^{\mathrm{TOT}}}10\sigma `$ level, increasing to about 10% at $`S_{^{\mathrm{TOT}}}5\sigma `$ \[Windhorst et al. 1990\].
The total weight of each source is then the multiple of the primary beam and resolution weights:
$$W=W_{\mathrm{pb}}\times W_{\mathrm{res}}.$$
(5)
These revised weights are given in the last column of Table 1, and are plotted in Fig. 1. This figure clearly shows how the weights rise exponentially near the survey flux density limit of 1 mJy. By applying a flux density limit of 2 mJy to the sample, the โcompletenessโ is increased significantly from 56% to 76% โ here defining the โcompletenessโ as the ratio of the number of sources actually detected to the total weighted number of sources. There are 9 sources with 1 mJy $`S_\nu <2`$ mJy that are excluded, with an average weight of 5 (a โcompletenessโ of 20%). Being the faintest objects, they have the largest errors and their effect on the sample is greatly exaggerated by the high weights. Therefore an a posteriori โ2-mJy sampleโ is defined, consisting of those 63 sources with $`S_{1.4}2`$ mJy. This will used in preference to the full Hercules sample in some of the statistical studies that follow in this and subsequent papers. It must be emphasized that following application of these weights, the incompleteness of the sample is fully corrected for and so it is, in effect, complete.
### 2.3 Radio properties of the Hercules sample
The radio data for the Hercules sample are summarized in Table 1. The coordinates are from the VLA 1.4 GHz observations where available \[Oort et al. 1987, Oort 1988\], otherwise the WSRT co-ordinates are used \[Windhorst et al. 1984a\]. The 1.4 GHz and 0.6 GHz fluxes ($`S_{1.4}`$ and $`S_{0.6}`$ respectively) and the resulting spectral indices ($`\alpha _{0.6}^{1.4}`$, where $`S_\nu \nu ^\alpha `$) are from the WSRT data \[Windhorst 1984, Windhorst et al. 1984a\]. The radio morphological data are taken from Oort et al. (1987) where available, otherwise the WSRT data are shown. Column 6 ($``$) notes whether the source is unresolved (U), resolved (R) or extended (E). An unresolved source consists of one component that is not distinguishably larger than the WSRT beam; a resolved source consists of one component whose extent is larger than the beam; and a source is classified as extended if it is clearly resolved into two (the classical double) or more components. Column 7 is the largest angular size ($`\psi `$) of the source in arcseconds, or an upper limit to $`\psi `$ for unresolved sources. The position angle ($`\varphi `$) of the largest angular size is given in degrees east of north. The weights ($`W`$) have been recalculated as discussed in the previous section and are given in the final column.
The flux density distribution of all 72 sources is presented as a histogram in Fig. 2. The median flux is 3.0 mJy, and for the 2-mJy sample the median is 6.6 mJy. The dashed line in the figure is the predicted distribution based on a 5<sup>th</sup>-order polynomial fit to the 1.4 GHz differential source counts of Windhorst & Waddington (2000). The fit was based on a sample of 10575 sources drawn from a large number of recent radio surveys. The overall agreement between data and model is indicative that the Hercules sample being used here is a fair representation of the full LBDS and indeed other surveys. We note that there is a small excess of bright sources at $`S_\nu >80`$ mJy, although, given the small area of the Hercules field, this may not be significant.
In Fig. 3 the distribution of 1.4โ0.6 GHz (21โ50 cm) spectral indices is plotted, for both the full Hercules sample and the 2-mJy sample (taking $`S_\nu \nu ^\alpha `$). Not all sources were detected at 0.6 GHz and thus those have only upper limits to their spectral index (indicated by the hatched histogram). The median spectral index of those sources with detections at both frequencies (i.e. excluding upper limits) is $`\alpha _{\mathrm{med}}0.8`$ for both samples. Taking into account the 0.6 GHz flux limits gives lower and upper limits to the median spectral index itself: $`0.05<\alpha _{\mathrm{med}}<0.74`$ for the full sample, and $`0.74<\alpha _{\mathrm{med}}<0.79`$ for the 2-mJy sample. Given the relatively few sources without spectral indices in the 2-mJy sample, the true $`\alpha _{\mathrm{med}}`$ is well constrained and we will adopt a value of 0.77 in subsequent work. Nearly half the full Hercules sample is without spectral index information, thus the small lower limit to $`\alpha _{\mathrm{med}}`$ above. However, we do not expect the true value to differ significantly from that of the 2-mJy sample, given that Donnelly et al. (1987) found that the median spectral index in the Lynx-2 area of the LBDS remains at $`\alpha _{\mathrm{med}}0.75`$ down to $`S_{1.4}0.25`$ mJy, when selected at 1.4 GHz.
## 3 Optical identifications and photometry
Twenty-five of the seventy-two sources in Hercules remained unidentified on the multi-colour Mayall photographic plates. These deep, sky-limited plates reached limiting magnitudes of $`U^+<23.3`$ mag, $`J^+<23.7`$ mag, $`F^+<22.7`$ mag and $`N^+<21.1`$ mag, roughly corresponding to a giant elliptical galaxy at $`z1`$ \[Windhorst et al. 1984b, Kron et al. 1985\]. Thus the unidentified sources were potentially the most important for investigating the high-redshift evolution of the faint radio source population. In order to complete the survey, deep CCD images were obtained of the unidentified radio sources.
### 3.1 Observations with the Palomar 200-inch
The Hercules field was observed on the 200-inch Hale telescope at Palomar Observatory between 1984 and 1988 (Table 4), using the 4-Shooter CCD camera \[Gunn et al. 1987\]. This camera consists of four simultaneously exposed $`800\times 800`$ pixel Texas Instruments CCDs, covering a nearly contiguous field of $`9\mathrm{}\times 9\mathrm{}`$, at a scale of 0.332 arcsec pixel<sup>-1</sup>. The total field of view is split by a reflecting pyramid into four separate beams which go to one each of the detectors. Poor reflections at the edges of the pyramid results in a strip of $`10`$โณ of sky between the CCDs which is not exposed, but $`>`$95% of the field is imaged by the detectors. One simultaneous exposure of the four CCDs is defined as a frame.
Sixteen separate pointings of the telescope were used in order to observe all of the unidentified radio sources (Fig. 4). Multiple observations were made through Gunn $`g`$, $`r`$ and $`i`$ filters over the six runs. Photometric standard stars were observed each night, normally at the start, middle and end of observing. Many of the nights were not photometric, though in almost all cases at least one photometric frame was obtained in each filter for each source. During the 1984 run, a problem with the 4-Shooter dewar resulted in liquid nitrogen coolant being spilled onto the pyramid thus obscuring much of the frame. Although taken in otherwise photometric conditions, these images were only occasionally used in the analysis, in cases where the target had not been observed at a later date and where the radio source was not affected by the spilt liquid nitrogen. The May 1987 run was used for spectroscopy and also an attempt to observe 53W043 by occulting a very bright star in the field that obscured the radio source. The July 1987 run was primarily spectroscopic data but a few images were also obtained.
### 3.2 Data reduction methods
The images were processed in a standard manner \[Neuschaefer 1992, Neuschaefer & Windhorst 1995\], but with some important differences. Examination of several bias frames confirmed that there was no two-dimensional structure to the bias (the mean counts in a $`10\times 10`$ pixel box varied by $`<0.3`$% across the image). For each exposure a first order spline was fitted to the 16-column bias strip and subtracted from across the image. The next step is normally to remove the dark current, however, the 4-Shooter CCDs have a very low dark current, which the observers considered to be negligible. Tests on the two best dark frames showed that the counts were comparable to the noise in the bias frames for exposures of about 1000 s (corresponding to a dark current of $`<10^3`$ e<sup>-</sup> s<sup>-1</sup> pixel<sup>-1</sup>), thus they were not used in the processing.
Flat-field images of the inside of the telescope dome were made at the beginning and end of each night. These images were the average of eight exposures of 7โ10 seconds each for $`g`$ and 2โ4 seconds each for $`r`$ and $`i`$. For each observing run, the mean of all the flat-field images was calculated, after scaling and rejection of bad pixels, to give a single flat-field for the whole of the run. The mean value of all the pixels in all four CCDs of the flat-field frame was calculated, and each image divided by this single value. Scaling in this way corrected for the global sensitivity differences between the four CCDs, assuming that they had been uniformly illuminated.
In some of the $`g`$ and $`r`$ images and in almost all the $`i`$ images, large-scale gradients remained across the CCDs, following division by the appropriate flat-field. Neuschaefer & Windhorst (1995) determined that the features in $`g`$ and $`r`$ were due to scattered light rather than sensitivity variations in the CCDs. Such global gradients were not important, as variations in the background of each target were removed at the photometry stage. The features in the $`i`$-band images posed a more serious problem, as they appeared to have two distinct sources. First, there were large-scale gradients across the images of a few percent of the sky background (larger than in $`g`$ and $`r`$), similar in appearance to the flat-field pattern itself. Rather than being scattered light, it is suggested that this was due to the different spectral characteristics of the night sky compared with the lamp used to illuminate the flat-fields, which was supported by the observation that the problem was most significant near twilight when the sky was changing most rapidly. This was then an extra multiplicative correction. The second problem with the $`i`$ images was that of fringing due to the strong night sky lines, on a smaller scale than the large gradients. This is an additive correction that scales with the brightness of the sky background.
Following Neuschaefer & Windhorst (1995), who actually only considered the fringing, an $`i`$-band โsuperflatโ was constructed. Most of the $`i`$-band data were taken in the June 1988 run, in which there were approximately 50 $`i`$-band exposures. Observations of the same region of sky were offset by 10โณโ20โณ and repeated no more than three or four times during the run, thus all the frames were effectively pointing at different parts of the sky. Following infrared imaging techniques, a superflat was constructed from the median of all 50 frames. This produced a high signal-to-noise picture of the sky flat-field, devoid of all sources. The flattened $`i`$-band images were divided by the superflat in order to remove the large-scale features. The fringing pattern was also removed by this process, although it is really an additive error. If fringing had been the dominant problem, then a scaled version of the superflat should have been subtracted from the images. This was done for several frames and the results compared with those obtained by dividing by the superflat. There was no visual difference between the results of the two methods, nor were they statistically different (both methods increased the sky noise by 0.5โ1.0%). For the $`i`$-band observations in earlier years there were not enough independent exposures to construct a superflat, however using the 1988 superflat often improved the images from these earlier runs.
Finally, the observations from different years were combined to produce an image of each radio source in the $`g`$, $`r`$ and $`i`$ filters. The images were registered using three or four unsaturated stars โ most images only required a linear offset, although a 0$`\stackrel{}{.}`$7 rotation was applied to the 1988 data. The images were then stacked, rejecting any remaining cosmic rays or bad pixels, in those cases where there were sufficient exposures to make this possible.
### 3.3 Optical identifications
The astrometry was performed on the $`r`$-band images in general. Occasionally the radio source was not visible in $`r`$, in which case either the $`g`$ or $`i`$ image was used. The Cambridge APM Guide Star Catalogue \[Lewis & Irwin 1997\] was used to find 10โ20 secondary standards in each image of interest. This catalogue is based on digitised scans of the first epoch Palomar Observatory Sky Survey, which has an internal accuracy of 0$`\stackrel{}{.}`$1โ0$`\stackrel{}{.}`$2, and is accurate to 0$`\stackrel{}{.}`$5 externally. A six-coefficient linear model was fitted to the 4-Shooter images, yielding root-mean-square residuals of $`0\stackrel{}{.}5\pm 0\stackrel{}{.}1`$.
The 1.4 GHz radio coordinates from Table 1 were converted to pixel coordinates on the 4-Shooter images. The errors in the radio positions are typically 0$`\stackrel{}{.}`$5โ1$`\stackrel{}{.}`$0. Combining these in quadrature with the optical astrometry errors defines a circle about the predicted coordinates of the source, with a 1-$`\sigma `$ radius of 0$`\stackrel{}{.}`$7โ1$`\stackrel{}{.}`$2, within which $`68`$% of the optical counterparts to the radio sources are expected to be found. Strictly, the errors in right ascension and declination are not equal and the search area should be an ellipse \[Windhorst et al. 1984b\], however with an ellipticity of typically 0.1โ0.2 the assumption of symmetrical errors was adequate for all identifications. Given the faintness of many of the sources, an optical identification was considered to be real only if the source was detected in at least two passbands. This requirement had the potential to exclude very red or very blue objects, but in practice no candidate identification was rejected on this basis.
A total of 47 sources were observed, of which 25 had not been identified on the Mayall photographic plates. Of these sources, new optical identifications for 19 are proposed here, 3 have been published elsewhere (53W002, 53W069, 53W091) and only 3 sources remain unidentified (53W037, 53W043, 53W087). Of the new identifications, 53W054 which was previously classified as a two-component radio source has now been reclassified as two independent objects with solid identifications (53W054A & 53W054B). Two sources have new identifications compared with those in Windhorst et al. (1984b): 53W051 has a close optical counterpart where it was previously associated with a source offset along the radio axis; and 53W022 is identified with a fainter source 5โณ to the north-west of the original candidate. The coordinates of all the identifications are given in Table 5 and the 4-Shooter images are presented in Fig. 5.
Three radio sources were not identified either on the photographic plates or in the 4-Shooter CCD observations, and a few comments can be made concerning their nature. The first of these, 53W043, is located 5โณ from a $`R=14`$ mag star, and lies within the wings of the starโs point spread function (PSF) in the 4-Shooter data. The PSF of the photographic plates is smaller than that of the CCDs, giving an upper limit to the brightness of the optical counterpart of $`F^+>23`$ mag. Counterparts to 53W037 and 53W087 were not identified on good 4-Shooter images, even after both smoothing the images and stacking the $`g`$, $`r`$ & $`i`$ images together. For both sources, an upper limit of $`r>25`$ mag was calculated. Neither source was detected in the infrared (see ยง4 below), with limits of $`K>20.7`$ mag for 53W037 and $`K>19.3`$ mag for 53W087. Both these objects are extended, steep-spectrum radio sources, suggesting that they may be classical radio galaxies. If so, then the faintness of their host galaxies suggests that they are either at very high redshift, or are obscured by dust, or are dusty high-redshift sources \[Waddington et al. 1999a\].
### 3.4 Photometry
The photometric calibration of the 4-Shooter images was done in two stages, due to the non-photometric conditions of a significant number of the observing nights. For most of the sources, at least one exposure had been taken on a photometric night which could therefore be calibrated accurately. This calibrated exposure was then used to bootstrap a zero-point for the non-photometric mosaic of all the exposures.
The apparent magnitude of a source ($`m`$) is defined by
$$m=m_{\mathrm{inst}}+k\overline{\mathrm{sec}z}+C(gr)+z_\mathrm{p}$$
(6)
where $`m_{\mathrm{inst}}`$ is the instrumental magnitude, $`k`$ is the extinction coefficient for a given filter, $`\overline{\mathrm{sec}z}`$ is the time-averaged airmass, $`C`$ is the colour coefficient for the filter, $`(gr)`$ is the instrumental colour of the source, and $`z_\mathrm{p}`$ is the zero-point. Higher order terms are not included due to the limited number of standard star observations available.
For each night of the observing runs, there were between three and eight observations of Gunn standard stars made. These were taken throughout the night at different airmasses and in all three filters. To avoid saturation, the standard stars were observed with the telescope out of focus. Instrumental magnitudes in a 30โณโ40โณ circular aperture were measured for each star, on every night that the log sheets recorded as being photometric. True apparent magnitudes ($`m_0`$) were taken from a list of both published and unpublished standards available at the telescope \[Thuan & Gunn 1976, Wade et al. 1979, Kent 1985\]. Equation 6 was then fitted to the data ($`m_0m_{\mathrm{inst}}`$, $`\mathrm{sec}z`$, $`gr`$) using a linear least squares method. This gave best-fitting values for $`k`$, $`z_p`$ and $`C`$ for each night. The colour coefficients are not expected to change from night to night and the fitted values were indeed consistent, thus a weighted average over all the nights in a run was used. A second least squares fit was then performed with $`C`$ fixed at its average value, and values for $`k`$ and $`z_\mathrm{p}`$ were found for each photometric night. For the non-photometric mosaics, the extinction and colour coefficients were taken as the mean of the values over all the runs (Table 8). Note that these values are consistent with those assumed by Neuschaefer & Windhorst (1995), who used the same instrument and telescope, over a similar period of time (1984โ1988).
For each photometric image containing a radio source, the automated image detection program pisa \[Draper & Eaton 1996\] was used to calibrate 20โ80 sources in the image. The search parameters were chosen such that a source was defined as having at least eight contiguous pixels with values greater than 5-$`\sigma `$ above the background. Such conservative parameters ensured that only good signal-to-noise sources were detected in the relatively short photometric exposure. The corresponding non-photometric, deep mosaic was then analysed in the same way with pisa. The mean difference in magnitude between each source, as observed in the mosaic and in the calibrated exposure, then allowed the zero-point of the non-photometric mosaic to be bootstrapped from the calibrated photometric image.
For five sources (53W010, 53W019, 53W020, 53W022, 53W027), only $`r`$-band observations from the 1984 run were available so no colour correction could be applied to them. In addition, only two standard star observations were available for each of the two nights, which was inadequate to determine both the extinction and zero-points. Thus, the mean extinction coefficient from Table 8 was adopted, and the zero-point calculated from the observed magnitudes of the standards. For five other sources (53W080, 53W081, 53W085, 53W086, 53W090), no photometric calibrations were available for the $`g`$ and $`r`$ images. The zero-points adopted for these mosaics were taken as the mean of the zero-points for all the other observations in the sample. This is reflected in the slightly larger errors for these sources in Table 5.
The sky background subtraction for each source was done using a method based upon the image reduction package cassandra by D. P. Schneider, adapted to run as an iraf task using the apphot package for the numerical work. A rectangular box (typically 10โณโ20โณ on a side) surrounding the source, and excluding any other objects, was cut out of the mosaic image. A plane was fitted to this box, excluding from the fit a circle centered on the source, and the sky was subtracted from the image. Photometry was performed using a circular aperture of 4$`\stackrel{}{.}`$0, 7$`\stackrel{}{.}`$5 or 10$`\stackrel{}{.}`$0 diameter (depending on the extent of the source), with the sky statistics determined over the remainder of the box. This enabled a large asymmetric sky aperture to be used, which is not possible with the basic apphot, in addition to providing a better measure of the sky background than is possible with an annulus around the object.
During testing of this method an error was found in the way that apphot calculates magnitude errors for faint objects. apphot makes the approximation $`\mathrm{ln}(1+\mathrm{\Delta })\mathrm{\Delta }`$ which for faint sources overestimates the error by as much as 0.1โ0.2 magnitudes. This was corrected by including the second, and for good measure the third, order terms in the expansion: $`\mathrm{ln}(1+\mathrm{\Delta })\mathrm{\Delta }\mathrm{\Delta }^2/2+\mathrm{\Delta }^3/3`$. The apparent magnitude was then calculated from equation 6 using the coefficients in Table 8 and the appropriate zero-points. The results are presented in Table 5. The errors quoted are the quadratic sum of the instrumental magnitude errors from apphot (shot-noise in the object aperture, sky-noise), and the errors in the determination of $`k`$, $`C`$ and $`z_\mathrm{p}`$.
In order to compare those radio sources with photographic magnitudes, $`J^+F^+N^+`$, to those sources with 4-Shooter Gunn magnitudes, $`gri`$, it was necessary to convert one photometric system to another \[Fukugita, Shimasaku & Ichikawa 1995\]. Windhorst et al. (1991) derived transformations from the Gunn system to the photographic system, and inverting these gives the required equations:
$`g`$ $`=`$ $`0.69J^++0.31F^+0.19`$ (7)
$`r`$ $`=`$ $`1.14F^+0.14J^++0.34`$ (8)
$`i`$ $`=`$ $`N^++0.75`$ (9)
For those sources without 4-Shooter observations, the optical magnitudes in Table 5 are photographic magnitudes from Kron et al. (1985) transformed to the Gunn system using these equations. An error of 0.3 magnitudes is assigned to each of these values, corresponding to the quadratic sum of the error in the transformations (0.1 mag) and the error in the photographic magnitudes (0.2โ0.3 mag). For the eighteen sources with both photographic and 4-Shooter Gunn photometry, the transformed photographic magnitudes are compared with the 4-Shooter data in Fig. 12. It can be seen that the transformations are in good agreement with the data for $`gri22`$โ23 mag; fainter than this, the $`J^+F^+N^+`$ magnitudes are close to the detection limits of the photographic plates and are less reliable.
## 4 Infrared photometry
Infrared data are crucial to understanding high-redshift galaxies. Beyond $`z1`$, optical observations sample the restframe ultraviolet emission of galaxies and the restframe optical emission has shifted into the near-infrared. Thus in order to properly compare sources at high-$`z`$ with those at low-$`z`$, the analysis must move to longer wavelengths. The $`K`$-band has long been recognized as a valuable window through which to investigate radio galaxies, and so we sought to obtain $`K`$ magnitudes for the complete Hercules sample.
### 4.1 Observations with the UK Infrared Telescope
Approximately half the sample has been observed during three major observing runs at the 3.8-m UK Infrared Telescope, Mauna Kea, Hawaii (Table 4). The first two runs (1992 & 1993) used the infrared camera IRCAM, a 62$`\times `$58-pixel InSb array with a scale of 0.62 arcsec pixel<sup>-1</sup>. All these observations were done in the $`K`$-band. The 1997 run used the upgraded camera IRCAM3, which had 256$`\times `$256 pixels with a scale of 0.286 arcsec pixel<sup>-1</sup>. Deep $`K`$ images were obtained for sources which had not previously been detected, together with $`H`$-band observations of several galaxies for which $`H`$ was considered important for constraining the redshift. Additional observations of some sources were made during associated projects and via the UKIRT Service Observing programme.
A standard jittering procedure was used to obtain a median-filtered sky flat-field simultaneously with the data. Nine 3-minute exposures were made, each offset by 8โณ (1992 & 1993) or 15โณ (1997) from the first position, resulting in a central area around the radio source with a maximal exposure time of 27 minutes, surrounded by a border of low signal-to-noise data. This was repeated two or three times for the faintest sources. Observations of faint standard stars were taken to calibrate the data.
### 4.2 Data reduction and calibration
For all three UKIRT runs, the image processing stage in the reduction was performed at the telescope by the observers. This consisted of the following steps: (i) a dark frame observed prior to the nine-exposure sequence was subtracted from each image; (ii) a normalised flat-field was constructed by taking the median of the nine exposures; (iii) each exposure was then divided by the flat-field; (iv) an extinction correction was applied; (v) the images were registered and finally combined into a mosaic after rejection of known bad pixels.
The standard star observations were reduced in the same manner, and the instrumental magnitude (corrected for extinction) calculated. Comparison with the known apparent magnitude then yielded a zero-point. Standards were observed regularly and the zero-points were found to vary by $`<0.05`$ mag throughout each photometric night. Colour coefficients are small in the infrared and were not determined.
Photometry was performed on each of the mosaics using the same method as for the optical data (ยง3.5), with the exception of the 1992 observations for which the sky background was determined from the mean of $`5`$ circular apertures placed randomly around the source. The results are presented in Table 5 and Fig. 5. In addition to the 36 new observations described here, Table 5 contains infrared data published by Thuan et al. (1984) for sixteen sources, unpublished data from G. Neugebauer, P. Katgert et al. (private communications) for a further twenty sources, and data on five other galaxies referenced in the notes to the table.
## 5 Discussion of the multi-colour photometry
The results of the optical and infrared photometry are discussed in this section, and compared with both earlier results from the LBDS and with results from other radio surveys. Discussion of redshift-dependent properties will be presented in Paper II, following analysis of the new spectroscopic data.
### 5.1 Notes on individual sources
Before investigating the photometric properties of the sample as a whole, the optical, infrared & radio observations of some individual sources are worth noting.
53W002 โ This $`z=2.390`$ radio galaxy has been the subject of much study (see ยง2.1). Comparing the optical and infrared colours recorded in Windhorst et al. (1991) with those here, shows that they are fully consistent within the errors, although there is some discrepancy between the optical magnitudes. This may be explained by noting that Windhorst et al. use โtotalโ magnitudes which are brighter than the 4โณ aperture magnitudes in Table 5 โ indeed, application of the same method to these data produces consistent results. It is important to note that the colours of faint sources are not significantly affected by aperture size so long as the aperture is larger than the seeing disk, which it is in all cases here.
53W011 & 53W013 โ There was originally some concern that these two sources may have been confused with the grating rings in the WSRT observations, however both have optical counterparts and so are retained in the sample.
53W014 โ Optical counterpart is at 3.1-$`\sigma `$ from the radio position. A second object, 5โณ to the north-east, at 3.7-$`\sigma `$, may also be a possible identification.
53W022 โ New identification compared with Windhorst et al. (1984b). This object looks as if it could be physically associated with the brighter old identification in the $`r`$-band. At $`K`$, the two objects are distinct and very similar in appearance.
53W032 โ A $`z=0.370`$ galaxy at the centroid of a double radio source. The 4-Shooter images reveal extended emission to the north-east, reminiscent of a tidal tail. Note also that component A is coincident with an $`r20.3`$ mag source, although component B does not have an optical counterpart.
53W035 & 53W089 โ These sources each appeared on two independent mosaics, enabling the repeatability of the photometry to be confirmed.
53W054A & 53W054B โ Originally classified as the lobes of a classical double radio source \[Windhorst et al. 1984a\], they are now considered to be two distinct sources. Both have solid identifications in the optical, and 53W054A is also identified in the infrared.
53W057 โ $`K`$ is a marginal (2-$`\sigma `$) detection at the north-east end of the elongated source in $`g`$. The brighter source at 6โณ to the west is also a possible identification, at 2-$`\sigma `$ from the radio position.
53W058 โ The radio source is in the arm of a bright spiral galaxy at $`z=0.034`$. The coordinates given in Table 5 are for the centre of the galaxy; the position of the radio emission is marked in Fig. 5.
53W067 โ The optical counterpart is 4$`\stackrel{}{.}`$8 (8-$`\sigma `$) from the radio position but has \[Oii\] and Ca H & K lines identified in the spectrum, at a redshift of 0.759 (Paper II). The fainter source to the south-east, originally proposed as the identification (at a separation of 1.2-$`\sigma `$), was identified as an M-type star from spectroscopy with the Keck telescope (Spinrad, Dey & Stern, 1997, private communication).
53W069 & 53W091 โ These two red sources at $`z1.5`$ have been extensively studied with both the Keck Telescope and the Hubble Space Telescope. Their rest-frame ultraviolet spectra are consistent with an evolved stellar population of ages 3โ4 Gyr \[Dey 1997, Dunlop et al. 1996, Dunlop 1999, Spinrad et al. 1997\], and their rest-frame optical morphologies are dominated by $`r^{1/4}`$ elliptical profiles \[Waddington et al. 2000b\].
### 5.2 General properties of the sample
The apparent magnitude distribution is of cosmological interest because it is closely related to the radio source redshift distribution. Broadly speaking, the fainter a source is (in optical/infrared wavebands), the more distant it is expected to be. Although complicated by evolutionary effects, in the absence of spectroscopic data the magnitude distribution has been an important tool in studying radio galaxies.
The optical data are shown in Fig. 13 and the infrared data in Fig. 14. In addition to the full Hercules distributions, the 2-mJy sample is shown as hatched histograms. The first issue that is apparent from these figures is the greater depth of the new observations. The limiting magnitudes of the Kron et al. (1985) photographic data correspond to $`g23.0`$ mag, $`r22.3`$ mag and $`i21.5`$ mag. The 4-Shooter data extend some 3โ4 magnitudes fainter in all passbands. The infrared $`K`$-band data are similarly about four magnitudes deeper than the Thuan et al. (1984) results. The $`griK`$ data are essential complete, approximately half the sources have $`H`$-band measurements and only one-third of the sample has been observed in $`J`$.
It is clear from Figs. 1314 that the number of galaxies per magnitude interval does not continue to increase down to the magnitude limit, but rather turns over at $`r22`$ mag. This immediately points to evolution of some form in the radio source population โ if the radio sources formed an homogeneous class of object present throughout the history of the universe, the numbers would continue to increase towards the magnitude limit. This turnover was hinted at in the $`N^+`$ magnitude distribution of the whole LBDS \[Windhorst et al. 1984b\] and is clearly confirmed for the Hercules subsample here. Excluding the 1 mJy $`S_{1.4}<2`$ mJy sources does not significantly change the shape of the magnitude distributions, but it does increase the median magnitude โ i.e. the faintest radio sources are not in general the faintest sources at optical/infrared wavelengths. This would suggest that the faintest radio sources are lower (radio) luminosity objects at lower redshift, rather than radio-powerful objects at high redshift. Some caution must be applied to this conclusion however, given that these sources are few in number (9) but have large weights ($``$2.5).
In Fig. 15, the LBDS $`K`$-band magnitude distribution is compared with the results of three other radio samples with near complete infrared data. The Parkes Selected Regions (PSR) data are principally taken from Dunlop et al. (1989). Only sources in their four-region subsample, for which the $`K`$-band data are essentially complete, are used (the 12<sup>h</sup> and 13<sup>h</sup> regions are poorly covered in the infrared). The 1-Jansky sample of Allington-Smith (1982) is a complete sample of 59 radio sources with 1 Jy $`<S_\nu <2`$ Jy at 408 MHz. Lilly, Longair & Allington-Smith (1985) presented infrared observations of 53 of these sources, which are plotted in the third panel of Fig. 15. Finally, a complete sample of 90 radio galaxies was selected from the 173-source 3CR catalogue of Laing, Riley & Longair (1983). Nearly complete $`K`$-band data was obtained for this 3CR sub-sample \[Lilly & Longair 1984\], which excludes all quasars, and is shown in the last panel of the figure. For the first three samples, both galaxies and quasars are shown. The LBDS panel also includes the 2-mJy sample for comparison.
As first pointed out by Dunlop et al. (1989), the 1-Jansky radio galaxies are biased towards larger $`K`$ magnitudes than the (radio) brighter 3CR galaxies. However, a similar shift is not seen between the 1-Jansky and PSR sources, a further factor of two fainter in flux. Dropping another factor of $``$200 in flux, the LBDS galaxies are seen here to have much the same distribution as the 1-Jansky and PSR galaxies. The median $`K`$ magnitudes for the four samples are 15.3 (3CR), 16.5 (1-Jy), 16.5 (PSR) and 16.8 (LBDS). To the extent that the $`K`$ magnitude of a powerful radio galaxy is a good estimator of its redshift, this result implies that although the 1-Jansky galaxies are consistent with being high-redshift counterparts of the 3CR galaxies, these same sources are not seen at even higher redshifts in the fainter samples. (Note that a median spectral index of $`\alpha =0.8`$ was used to bring the flux limits of the samples to a common frequency for this comparison.)
Although the number of sources is small, the same effect is not seen in the quasar distributions. The median $`K`$ magnitude of quasars in the fainter three samples increases with decreasing flux limit: 16.1 (1-Jy), 16.3 (PSR) and 16.8 (LBDS). This may suggest that the quasar population is not changing as rapidly as the radio galaxy population. It must be noted, however, that the decrease in restframe optical/infrared flux between the 1-Jansky and LBDS quasars (a factor of $`2`$) is certainly not comparable to the drop in radio flux (a factor of $`400`$) โ i.e. if the $`K`$ magnitude is a fair redshift estimator, the LBDS quasars must be less luminous in the radio than the 1-Jansky quasars.
Fig. 16 presents colourโmagnitude diagrams for the Hercules sample. To compare these results with those of Kron et al. (1985), note that the colours are related by $`(gr)(J^+F^+)0.5`$ and $`(ri)(F^+N^+)0.4`$ (from equations 79). It can be seen that the colours of the new identifications are comparable to those of the photographic identifications at $`r22`$ mag. There is some bias towards bluer $`gr`$ (and to a lesser extent, $`ri`$) colours at the fainter magnitudes. In particular, there is a deficit of faint ($`r>23`$ mag), red ($`gr>0.5`$, $`ri>1`$) radio sources compared with field galaxy surveys of comparable depth \[Neuschaefer 1992, Neuschaefer & Windhorst 1995\]. This is consistent with a continuation of the faint blue radio galaxy population that Kron et al. (1985) found at brighter magnitudes.
The sources classified as quasars (Q & Q?) tend to have bluer optical colours on average than the galaxies, as one would expect from the dominance of the AGN at ultraviolet/optical wavelengths in such sources. (Recall that the classification was based on morphology, not colour.) The โgalaxyโ class includes objects of unknown identity (?), which dominate at the faintest magnitudes. Figs. 16(a), (b) & (f) suggest that the bluest of these objects may in fact be quasars. Fig. 6 of Dunlop et al. (1989) shows that essentially all PSR sources with $`K15`$ and $`RK3`$ are quasars. Comparing that figure with Fig. 16(f) here, shows that the same cannot be said for the LBDS โ only one of these faint blue objects is of uncertain classification (53W068) and one is a probable quasar (53W036), while the other three are G (53W034, 53W067) and G? (53W027). The LBDS quasars have a much broader range of colours than those in the PSR, perhaps a consequence of their weaker AGN.
Figs. 16(c), (d) & (e) show the distribution of near-infrared colours with apparent magnitude. In considering these figures, it must be remembered how incomplete these data are (only one-third of the sample has $`J`$-band data, only one-half has been observed in the $`H`$-band). There appears to be a linear correlation between $`iJ`$ colour and $`i`$ magnitude in Fig. 16(c), however this may simply be due to the absence of faint $`J`$-band photometry โ at any given $`i`$ magnitude it is the reddest sources in the sample that would have been preferentially detected at $`J`$. The dispersion of the $`JH`$ colours (0.2 mag), defined as the standard deviation about the mean colour, is 2โ3 times smaller than the dispersion in the optical colours. This is probably due to the incomplete infrared data, but may also be a genuine property of the sample. The smaller scatter may indicate that the infrared emission is dominated by stellar light from the host galaxies, rather than emission from the AGN which tends to dominate at bluer wavelengths and can vary significantly from one source to another.
To summarize, the optical identification of the 72-source LBDS Hercules sample is essentially complete (only three sources remain unidentified). Approximately 85% of the sources have also been identified in the near-infrared $`K`$-band and a programme to obtain complete $`J`$\- and $`H`$-band data is ongoing. Paper II will present the results of work to measure redshifts for this sample from both direct spectroscopy and by using photometric estimation techniques. This is now the most comprehensive sample of radio-selected sources available at millijansky flux density levels, and is ideally suited to study the evolution of the 1.4 GHz radio luminosity function (for AGN) out to very high redshifts โ a topic to be addressed in forthcoming papers.
## Acknowledgments
We thank M. Oort and S. Anderson for contributing to the Palomar observations; T. Keck for assisting with the processing of the 4-Shooter data; and P. Katgert, G. Neugebauer, H. Spinrad and D. Stern for allowing us to include some unpublished data. The Hale Telescope at Palomar Observatory is owned and operated by the California Institute of Technology. The United Kingdom Infrared Telescope is operated by the Joint Astronomy Centre on behalf of the PPARC. This work was supported by a PPARC research studentship (to IW); and by NSF grants AST8821016 & AST9802963 and NASA grants GO-2405.01-87A, GO-5308.01-93A & GO-5985.01-94A from STScI under NASA contract NAS5-26555 (to RAW).
|
warning/0006/astro-ph0006382.html
|
ar5iv
|
text
|
# Spectroscopic confirmation of a cluster of galaxies at ๐ง=1 in the field of the gravitational lens MG2016+112 Based on observations made with the Nordic Optical Telescope, operated on the island of La Palma jointly by Denmark, Finland, Iceland, Norway, and Sweden, in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofisica de Canarias. Based on observations with the Canada-France-Hawaii Telescope at Mauna Kea, Hawaii, USA.
## 1 Introduction
Gravitational lensing is an efficient way to probe the mass content of the Universe. It acts on many scales, from dark stellar-mass objects to the largest massive structures Mellier (1999). Some of the most spectacular examples of this phenomenon are the multiple quasar systems already known for 20 years Walsh et al. (1979). However, in most cases simple lens models cannot reproduce the image geometry and flux ratios, unless an additional external component such as external shear is included in the modeling Keeton et al. (1997). Furthermore in the cases of large-separation multiple quasars, the mass-to-light ratio of the main galaxy lens is generally much higher than what is usually expected for a galaxy. As an example, the double quasar Q0957+561, with a separation of 6โณ, is well explained with a deflector corresponding to a massive and bright galaxy combined with the additional effect of a cluster of galaxies centered on it (Bernstein & Fisher 1999, and references therein). This cluster was indeed spectroscopically confirmed at the redshift of the main galaxy deflector Angonin-Willaime et al. (1994). In general, it is becoming evident that the likely explanation of the external shear necessary to understand the wide separation lensing configurations is the existence of a group/cluster near the line of sight of these multiple quasars. In particular, cluster mass distributions have been detected in several other wide-separation multiple quasars such as the so-called โCloverleafโ Kneib et al. (1998) and RX J0911+05 Burud et al. (1998); Kneib et al. (2000).
A mysterious system in this respect is the triple quasar MG2016+112 ($`z=3.26`$). It was discovered by Lawrence et al. Lawrence, Schneider, Schmidt, Bennett, Hewitt, Burke, Turner, & Gunn (1984) and studied in more detail by Schneider et al. Schneider, Lawrence, Schmidt, Gunn, Turner, Burke, & Dhawan (1985, 1986) in the optical and by Garret et al. Garrett, Muxlow, Patnaik, & Walsh (1994, 1996) in the radio. The third image named C presents a complex structure, suggesting some differential magnification of the source within each image. Deep imaging revealed the existence of a giant elliptical galaxy at $`z1.01`$ named D and located between the 3 images. It is identified as partly responsible for the multiple lensing of the quasar, although its mass is insufficient to explain the complex nature of the lensed images of MG2016+112. Detailed modeling requires another source of lensing mass, such as the mass of a rich cluster, centered on galaxy D Narasimha et al. (1987). In order to reveal the exact nature of the lens, deep X-ray searches have been attempted to probe hot intra-cluster gas. A positive X-ray detection of a cluster-like emission was obtained, centered on MG2016+112 Hattori et al. (1997). The X-ray characteristics (notably an emission line at $`3.49_{0.13}^{+0.15}`$ keV and the extended nature of the emission) indicate that AX J2019+1127 is an X-ray cluster of galaxies at $`z1`$. The X-ray temperature related to this emission gives $`k`$T = $`8.6_{3.0}^{+4.2}`$ keV and an X-ray luminosity L$`{}_{X}{}^{}=8.4_{1.7}^{+2.4}h_{50}^2`$ erg s<sup>-1</sup> in the 2โ10 keV band, quite compatible with the L<sub>X</sub>โT<sub>X</sub> relation for clusters of galaxies Markevitch (1998). The mass derived from the X-ray analysis is about $`3\times 10^{14}h_{50}^1`$ M within 500 $`h_{50}^1`$ kpc, a typical value for rich clusters of galaxies with velocity dispersion of $`1200`$ km s<sup>-1</sup> (assuming the observed $`\sigma T_X`$ relation, Girardi et al. 1996). A re-analysis of the HRI/ROSAT imaging Benรญtez et al. (1999) seems to slightly decrease the expected cluster mass, without rejecting its detection. In addition, deep near-IR imaging of the field was used to identify a โred sequenceโ in the color-magnitude diagramme, presumably due to early-type galaxies in the cluster.
Recently, Clowe et al. Clowe, Trentham, & Tonry (2000) revisited the Benรญtez et al. optical data, complemented with ultra-deep Keck R-band images. Again a sequence of cluster galaxies was clearly identified and its spatial distribution analyzed. More interesting is their weak lensing analysis of the faint galaxies from the R-band image, which seems to give a signal in the field. Its center appears offset about 1โฒ North of MG2016+112 itself and its mass is consistent with the mass inferred from X-ray data. The authors claim they can rule out a mass centered exactly on MG2016+112, at a $`2\sigma `$ level. Complementary data are clearly required to better quantify these preliminary results of relatively low significance.
In this paper, we present new spectroscopic observations in the field of this peculiar lens, aimed at identifying the lensing cluster. Observations and data reduction are described in Section 2, the results are presented in Section 3. Section 4 presents a preliminary analysis of the lensing cluster identified at $`z1`$. Section 5 gives some conclusions and prospects in the understanding of this object.
Throughout the paper, we use a Hubble constant of H$`{}_{0}{}^{}=50h_{50}`$ km s<sup>-1</sup> Mpc<sup>-1</sup>, with $`\mathrm{\Lambda }=0`$ and $`\mathrm{\Omega }_0=1`$. At the cluster redshift ($`z=1`$), 1โณ corresponds to 8.52 $`h_{50}^1`$ kpc and 1 $`h_{50}^1`$ Mpc covers 2โฒ.
## 2 Observations and data reduction
### 2.1 Photometry
The imaging data were obtained in R and I-bands in June 1996 at the Nordic Optical Telescope, using the HiRAC instrument with a LORAL 2k$`\times `$2k CCD, giving a pixel size of 0$`\stackrel{}{.}`$107. The data reduction was performed in the IRAF<sup>1</sup><sup>1</sup>1IRAF is distributed by NOAO, which are operated by AURA, Inc., under cooperative agreement with NSF. environment. Observations of M92 at various airmasses provided a calibration to the Landolt system Landolt (1992) based on CCD photometry of the field (L. Davis, private communication). The zero-point, color coefficient and extinction was then computed using the PHOTCAL package. In the relatively crowded field of MG2016+112 bright foreground stars with a significant PSF tail extending to many tens of arc-seconds hampered conventional photometry of the much fainter galaxies of interest. Moreover, PSF subtraction of point sources in the combined image proved difficult due to interpolation artifacts during the image mapping to a common coordinate system. Instead the bright unsaturated stars were removed by PSF subtraction on the individual images. To compensate for possible PSF variations in the field we experimented with both a constant and second order varying PSF. No significant differences were detected, however, and the constant PSF was therefore used. The spatial image transformations determined from the pre-PSF subtracted images were then applied and combined to get the final R and I images. The FWHM of unsaturated stars in these images were found to be 0$`\stackrel{}{.}`$75 and 0$`\stackrel{}{.}`$69 in the R and I-bands, respectively.
Galaxy photometry was computed using the MAG\_BEST estimate from SExtractor Bertin & Arnouts (1996) and transformed to the standard Landolt system using the derived photometric transformation. The photometric catalogue was not corrected for Galactic extinction although for this particular field, the correction could be as high as 0.5 magnitude, depending on the filter. Finally conversion to AB magnitudes was done using coefficients (c$`{}_{\mathrm{R}}{}^{}=0.169`$, c$`{}_{\mathrm{I}}{}^{}=0.432`$) given in Fukugita et al. Fukugita, Shimasaku, & Ichikawa (1995). A total of 7 R-band images ($`4500`$ sec) and 10 I-band images ($`6000`$ sec) were combined reaching 3$`\sigma `$ sky limits of $`25.4`$ and $`25.3`$ within the seeing disk (1.5$`\times `$FWHM), respectively. A significant fraction of faint (cluster member) galaxies are expected to remain undetected owing to the strong halos of saturated stars in the field and the fact that the photometry of the field is not extremely deep.
### 2.2 Spectroscopy
Spectroscopic data were acquired on the nights August 2โ4, 1997 at the Canada-France-Hawaii Telescope with the OSIS multi-object spectrograph Le Fรจvre et al. (1994). A tip-tilt correction is introduced thanks to the guiding on a selected star in the field. The seeing is then significantly improved ($`0\stackrel{}{.}5`$ measured on our images), increasing the detectivity of the instrument for faint object spectroscopy. We used the $`2048\times 2048`$ STIS2 thinned CCD, with pixels of 21 $`\mu `$m (or 0$`\stackrel{}{.}`$151 on the sky) covering a field of view of $`3\stackrel{}{.}6\times 3\stackrel{}{.}6`$. We also used the R150 grism, giving a useful spectroscopic signal from 4500 to 9000 ร
and a dispersion of 3.7ร
/pixel. The slits were 10โณ long and 1$`\stackrel{}{.}`$0 in width, giving a final instrumental resolution of 18ร
. Two masks were punched with 19 slits and 18 slits respectively, and observations were obtained for 21600 seconds for Mask1 (6 exposures) and 23400 seconds for Mask2 (7 exposures) in good weather conditions with seeing ranging from 0$`\stackrel{}{.}`$5 to 0$`\stackrel{}{.}`$8. The spectroscopic targets were selected from the photometric catalogue, with the following criteria: $`22<I_{AB}<23.3`$ and $`(RI)_{AB}>0.5`$, with a careful selection of resolved objects. These criteria were optimized to avoid a strong contamination of the sample by the faint stars crowding at this low galactic field, and to increase the chance of selecting high redshift galaxies.
Data were reduced with the MULTIRED package Le Fรจvre et al. (1995). Spectra were bias-subtracted, flat-fielded and extracted. The wavelength calibration gave typical internal errors of $`0.3`$ ร
. The spectra were finally rebinned to a dispersion of 6ร
/pixel, better matched to the resolution of the instrument. We encountered some difficulties with the flux calibration in the red end of the spectra because no order-separating filter was used. So contamination by the second order contribution was significant above 7500โ8000ร
, giving lower confidence to the flux calibration above this limit.
Finally, the redshift measurement of the spectra was done, using the RVSAO2.0 package Kurtz & Mink (1998) and correlating the observed spectra with a set of stellar and galactic templates. Because of the low S/N for most of the spectra, a visual check of the spectral identification was done carefully as well, in order to help the redshift measurement of the less significant spectra.
Among the 44 spectra extracted (in some of the 37 slits, more than 1 spectrum was extracted), 10 correspond to stars and 3 remain unidentified due to a poor S/N ratio, but these are most probably non-stellar. For the 31 remaining spectra the redshift identification is presented in Table 1 and in Fig. 1. The redshift distribution of these 31 objects is shown in Fig. 2. The sample is far from complete in magnitude so no conclusion can be drawn on the redshift distribution of the galaxies along this line of sight. An excess of galaxies around $`z=1`$ is, however, clearly seen and corresponds to the redshift of galaxy D already identified in Schneider et al. Schneider, Gunn, Turner, Lawrence, Hewitt, Schmidt, & Burke (1986). Two other concentrations are suspected which are also spatially coherent at $`z0.66`$ and $`z0.82`$, with a somewhat lower significance.
## 3 The lensed source MG2016+112
Most of our results on the multiple images of the radio source MG2016+112 are presented in Yamada et al. Yamada, Yamazaki, Hattori, Soucail, & Kneib (2000) who study in detail the emission-line properties of the images B and C. Note that this is the first spectrum of C observed in the optical (Fig. 3), after its identification as a third image by its Ly$`\alpha `$ emission Schneider et al. (1986) and its radio emission Garrett et al. (1994). The higher resolution of our spectrograph compared to the spectrum of B presented in Schneider et al. Schneider, Lawrence, Schmidt, Gunn, Turner, Burke, & Dhawan (1985) allows to resolve the main emission lines and then to study the physics of the emitting regions in the source.
Schneider et al. Schneider, Gunn, Turner, Lawrence, Hewitt, Schmidt, & Burke (1986) claimed that 2 objects, labeled A1 and B1, detected close to the triple radio-source, were at the same redshift as the lensed source. They identified them from a narrow-band image of the field, centered on Ly$`\alpha `$ redshifted to 3.26. A close look at our deep I and R images confirms the detection of B1 but not A1. B1 was positioned in one slit of Mask 2 and its spectrum (Fig. 3) confirms its redshift at $`z=3.269`$, similar but different from MG2016+112. Note however that the emission-line properties of B1 are quite different from those of B, with only Ly$`\alpha `$ in emission. SiIV and CIV may be detected in absorption but with a lower confidence. If true, these lines are representative of a young stellar component in the source, rather than an active nucleus.
## 4 The lensing cluster at $`z1`$
In our spectroscopic sample, 9 objects have a redshift $`z1`$, with 8 new spectroscopic identifications (Fig. 4). This confirms that a coherent gravitational structure is associated with object D, previously suspected as the main lens Hattori et al. (1997). Not surprisingly, these objects are among the reddest ones in Table 1. This justifies a posteriori the color selection we introduced in the mask production. Moreover, if one concentrates on the objects selected in a 0.5 $`h_{50}^1`$ Mpc radius around galaxy D, we note that all galaxies are at $`z1`$ except one at $`z=0.52`$. As there are not many other red galaxies in this region, we may have identified most of the brighter cluster galaxies. In this respect deep IR imaging should help to identify fainter cluster members, because the foreground contamination is less important in the near-IR.
From multicolor photometry only, Benรญtez et al. Benรญtez, Broadhurst, Rosati, Courbin, Squires, Lidman, & Magain (1999) followed by Clowe et al. Clowe, Trentham, & Tonry (2000) attempted to isolate cluster galaxies from their colors in the red and near-infrared. Of course this selection procedure favors old-type galaxies and avoids strong star-forming ones which are more likely mixed with field galaxies. We compared their sample with our spectroscopically confirmed cluster members and found a good agreement with their selected objects. In particular we confirm the cluster membership of objects B5 ($``$ M1\_9, $`z=1.001`$), B6 ($``$ M1\_13, $`z=1.011`$) and B2 ($``$ M2\_11, $`z=1.000`$), while we find a lower redshift $`z=0.8231`$ for object labeled # 3 in Clowe et al. Clowe, Trentham, & Tonry (2000). Note that their identification was done after our preliminary results presented in the CFHT Inofrmation Bulletin (report # 38, 1998), so their sequence may not be quite independent from our first results.
Among the 9 galaxies at $`z1`$, 4 galaxies (+ 2 more uncertain ones) show the \[OII\] 3727ร
line in emission, indicating active star formation in these galaxies. This may be a sign of the relative youth of this cluster or sign of interaction between the galaxies and the intra-cluster medium. Note that similar trends were observed in a few other high redshifts clusters Postman et al. (1998). To address in more detail the question of the evolution of galaxies and their spectral content in high redshift galaxies, a large sample of cluster members should be analyzed spectroscopically as well as through their optical and near-IR colors, including morphological information. This kind of data is now within the reach of 8โ10m class telescopes equiped with optical and near-IR instruments.
From a dynamical point of view, we can derive a rough estimate of the velocity dispersion, although with small numbers it is difficult to measure accurate properties. If we keep the total sample of 9 galaxies, we find a mean redshift of $`\overline{z}=0.995`$ and a velocity dispersion of $`\sigma _{los}=2510_{450}^{+960}`$ km s<sup>-1</sup>. The maximum difference is $`\mathrm{\Delta }v=1/2(v_{max}v_{min})=\pm 3500`$ km s<sup>-1</sup>. These values are rather unrealistic for a well defined dynamical structure. But among our 9 galaxies, 3 are isolated around a redshift $`z=0.97`$, while the 6 others are spread around $`z=1`$. We suspect the first 3 galaxies may belong to a small structure such as a group, spatially and dynamically close to the main one. But we are well aware that these assumptions are quite unsecure because of the small numbers. Considering then the 6 galaxies in the redshift range $`[1.0;1.012]`$ we find a mean redshift of $`\overline{z}=1.005`$, a velocity dispersion of $`\sigma _{los}=771_{160}^{+430}`$ km s<sup>-1</sup> and a velocity range: $`\mathrm{\Delta }v=\pm 900`$ km s<sup>-1</sup>. These values make more sense in view of the X-ray observations, although the error bars are still quite large.
Using the virial theorem, we can estimate the mass of this system within the harmonic radius $`R_h`$ defined as Nolthenius & White (1987):
$$R_h=D_A(\overline{z})\frac{\pi }{2}\frac{N_m(N_m1)}{2}\left(\mathrm{\Sigma }_i\mathrm{\Sigma }_{j>i}\theta _{ij}^1\right)^1,$$
(1)
where $`\theta _{ij}`$ is the angular distance between galaxies $`i`$ and $`j`$, $`N_m`$ is the number of cluster members, and $`D_A(\overline{z})`$ is the angular diameter distance at the mean cluster redshift $`\overline{z}`$. The cluster virial mass can then be estimated as
$$M=\frac{6\sigma ^2R_h}{G}.$$
(2)
With our sample of 6 galaxies, we find an harmonic radius of $`R_h=348h_{50}^1`$ kpc and a virial mass of $`M=2.8_{1.0}^{+4.0}10^{14}h_{50}^1`$ M. The error bars quoted in the virial mass correspond to the errors in the velocity dispersion measure only. They do not include uncertainties due to the harmonic radius although they may be significant, because of our very small sample of cluster members. As quoted by Carlberg et al. Carlberg, Yee, Ellingson, Abraham, Gravel, Morris, & Pritchet (1996), the use of the harmonic radius may underestimate the virial mass determination because $`R_h`$ is highly sensitive to close pairs and rather noisy.
We can also compare our dynamical mass with the galaxy distribution to get an estimate of the mass-to-light ratio of the structure. Our own measure of the luminosity of the 6 cluster members, corrected the same was as Benรญtez et al. Benรญtez, Broadhurst, Rosati, Courbin, Squires, Lidman, & Magain (1999) \[k-correction of 3.42 in V and 1.12 in I, average color index $`VI=3.2`$ for the galaxies, average extinction correction $`A_I=0.44`$ magnitude\] gives a total luminosity of $`L_V^{gal}=7.6\times 10^{11}h_{50}^2L_V`$ for these galaxies. It is about 50% higher than the luminosity of their 9 cluster member candidates and 60% smaller than their extrapolated total magnitude. A more accurate luminosity function determination is necessary for a correct estimate of the total luminosity, and the M/L ratio. Following the reasonable total luminosity function of the cluster galaxies estimated by Benรญtez et al. Benรญtez, Broadhurst, Rosati, Courbin, Squires, Lidman, & Magain (1999) ($`L_V^{all}=1.3\times 10^{12}h_{50}^2L_V`$) we find a mass-to-light ratio of
$$(M/L_V)=215_{77}^{+308}h_{50}(M/L_V)_{}$$
This value can be compared with two other estimates proposed for this cluster. From the X-ray emission and the measurements of Hattori et al. Hattori, Ikebe, Asaoka, Takeshima, Bรถhringer, Mihara, Neumann, Schindler, Tsuru, & Tamura (1997) we find a value $`M_X/L_V=224h_{50}(M/L_V)_{}`$ within a radius of 400 $`h_{50}^1`$ kpc. Note that an error of a factor two occurs in the estimate of $`M_X/L_V`$ in Benรญtez et al. Benรญtez, Broadhurst, Rosati, Courbin, Squires, Lidman, & Magain (1999), due to an error in the conversion between $`h_{50}`$ and $`h_{100}`$ units. From an extrapolation of their lens model they also find a lensing $`M/L_V`$, within a radius of 400 $`h_{50}^1`$ kpc of 186 $`h_{50}(M/L_V)_{}`$. Our determination falls within the same range as these two estimates. All of them are based on the cluster luminosity estimated by Benรญtez et al. which is most probably underestimated because of the difficulties in selecting cluster members in the optical in this low galactic latitude field. In addition, no evolution corrections are included, which at such high redshift would increase the $`M/L`$ ratio, typically by a factor of 2 Smail et al. (1997). Better estimates will come from near-IR imaging where both effects are strongly reduced. In all cases, the $`M/L`$ value proposed for MG2016+112 is slightly higher than other accepted values found in other high redshift clusters: Smail et al. Smail, Ellis, Dressler, Couch, Oemler, Sharples, & Butcher (1997) find an average $`M/L_V=90h_{50}(M/L_V)_{}`$ within 800 $`h_{50}^1`$ kpc for their sample of clusters in the redshift range \[0.2โ0.5\], while Carlberg et al. Carlberg, Yee, Ellingson, Abraham, Gravel, Morris, & Pritchet (1996) have a median value of $`M/L_r=143h_{50}(M/L_V)_{}`$ for the CNOC sample. For a more distant cluster ($`z1`$), Deltorn et al. Deltorn, Le Fรจvre, Crampton, & Dickinson (1997) find a slightly higher value of $`M/L_B=200h_{50}(M/L_B)_{}`$ within a radius of 400 $`h_{50}^1`$ kpc from 11 cluster members. In any case, MG2016+112 is not a โdark clusterโ anymore with a $`M/L`$ ratio as high as 1000 as suspected initially when most of the light was concentrated on galaxy D.
## 5 Conclusions
We have presented a photometric and spectroscopic survey of the field around the triple lens MG2016+112. Our observations confirm the identification of a massive cluster at $`\overline{z}=1.005`$ Hattori et al. (1997). With the identification of 6 secure cluster members we have estimated the velocity dispersion $`\sigma _{los}800`$km s<sup>-1</sup> and used this to infer the virial mass associated with the cluster. Although we are well aware of the limitations due to the small number statistics of our sample in the analysis of the cluster mass, we also tried to compare with other estimates such as the mass inferred from the X-ray analysis or a weak lensing one. All the M/L ratios derived that way are consistent with each other. The value found ($`M/L=215_{77}^{+308}h_{50}(M/L)_{}`$ for the dynamical analysis) is also consistent with what is found in other high redshift clusters, although we cannot yet deduce any evolution in this ratio. But clearly the cluster around MG2016+112 is not a dark cluster anymore.
To overcome most of the observational difficulties due to the low galactic latitude of the field, we underline the need of deep near-infrared imaging to increase the density contrast of $`z=1`$ galaxies clustered around the lensed source and to reduce the uncertainties in the luminosity function of these galaxies. A deeper spectroscopic follow-up with a spectrograph on a 8โ10m telescope will also increase the identification of cluster members and a better dynamical analysis. Furthermore, deep X-ray observations with the 2 new X-ray satellites, Chandra and XMM-Newton, will also increase the accuracy in the cluster X-ray emission (exact location of the center, good determination of the mass profile, better measure of the temperature and Fe abundance, etc.), leading to a more accurate cluster mass measure.
As the number of clusters of galaxies at redshift $`z>1`$ is still small it is valuable to focus on their details to better understand the physical properties and the rate of evolution, necessary to relate them to the local distribution of clusters of galaxies. Large separation quasars are probably a good alternative way to detect high redshift structures.
###### Acknowledgements.
We wish to acknowledge fruitful discussion with D. Clowe, Y. Mellier and P. Schneider. JPK acknowledges support from CNRS, JH from the Danish Natural Science Research Council (SNF).
|
warning/0006/astro-ph0006349.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The observation of Ultra High Energy Cosmic Ray (UHECR) events with energies above $`10^{20}eV`$ is one of the most intriguing puzzles of modern physics. The point is that if UHECR come from large distances their energy dramatically decreases due to interactions with background fields. For example, nucleons of energies above $`4\times 10^{19}eV`$ start to interact with the Cosmic Microwave Background (CMB) and lose their energy due to pion photoproduction (the so-called Greisen-Zatsepin-Kuzmin (GZK) effect ). As a result, nucleons can not travel to an observer from distances larger than $`R_{GZK}50Mpc`$. Heavy nuclei travel even much less distances of about few Mpc because they photodisintegrated by CMB. UHE electrons lose their energy due to synchrotron radiation in extragalactic magnetic fields. From other side, it is hard to accelerate protons, electrons and heavy nuclei up to such energies even in the most powerful astrophysical objects such as radio galaxies and active galactic nuclei, but even if it is possible, usually such objects are situated far away from our galaxy on distances larger than $`R_{GZK}`$. This means that energy of a UHE particle will drop below the GZK threshold unless its energy is larger than observable one on several orders of magnitude.
It was suggested several ways to avoid this difficulty, but all of them require either the introduction of a new physics well beyond the Standard Model or extreme astrophysics.
One possible way to avoid GZK cutoff is to introduce a new particle, which interactions with CMB are suppressed in compare with the proton of the same energy. The possible examples of such particles are light long-leaving gluino-containing baryons or axion-like particles .
Another possibility is to stand with Standard Model particles but change physics laws. For example, protons and neutrons still can be UHECR with energies larger then cutoff if one slightly violate Lorentz invariance .
One more way avoid a GZK cutoff is to produce UHE protons and photons inside of GZK volume from hidden sources. Photons with extremely high energies $`E>10^{22}`$ eV have attenuation length larger then 100 Mpc, if Extragalactic Magnetic Fields (EGMFs) are small $`B_{\mathrm{EGMF}}<10^{11}`$ G. Such photons can be the sources of the secondary UHE photons inside of GZK sphere . Another possible source of both UHE protons and photons are neutrinos with $`E>10^{22}`$ eV, which could interact with background neutrinos and produce secondary particles through $`Z^0`$ resonance . Both those scenarios require extreme astrophysical sources which should accelerate primary photons up to energies $`E>10^{23}`$ eV.
The other way to explain UHE events above the GZK cutoff is to create UHE protons, leptons and photons in decays of supermassive X-particles inside of the GZK sphere. This is one of the so-called โtop-downโ (TD) scenarios. Depending upon the particular scenario, decaying X-particles could populate the Galaxy or can decay on cosmological distances . In the latter case the important role is played by their decay secondariesโ energy losses during propagation through the space. The sources of the massive X-particles themselves might be topological defects such as cosmic strings or magnetic monopoles that could be produced in the early Universe during symmetry-breaking phase transitions envisaged in Grand Unified Theories . In an inflationary early Universe, the relevant topological defects could be formed at a phase transition at the end of inflation. Other possibility is to produce superheavy long-living particles thermally during reheating epoch of the Universe or from vacuum fluctuations during or after inflation . For a review of TD models see, for example .
Important information about cosmic ray sources give the measurement of the UHECR arrival directions. All experiments see uniform distribution of the UHECR over the sky. This leads to the conclusion that most probably UHECR come from the extragalactic sources. Besides, AGASA experiment see small scale clusters . Those clusters can come from point-like sources . Moreover, the correlations of UHECR with BL Lacertae was found . If those small scale clusters and especially correlations with BL Lacs will confirmed in the future, TD models will be excluded.
The main goal of this work is a self-consistent consideration of the galactic and extragalactic contributions to spectra of UHECR from TD models. There are two major possibilities: if sources of UHECR are present in our Galaxy or not. In the first case simple estimates show that contribution of the Galaxy component is about 100 times larger than extragalactic component and hence one might take into account the Galactic contribution only. Our numerical calculations, based on detailed modeling of UHECRs propagation process give even more strict limitations on the ratio of the extragalactic component of UHECRs to the galactic one.
The main observational signature allowing to distinguish this kind of models is a quite significant anisotropy in all the UHE particles spectra which is in fact simply due to the non-central position of the Sun in the Galaxy at the distance $`8.5kpc`$ from the Galaxy center. In what follows we present our results for various models of UHECR source distributions in the Galaxy.
In the case there is no sources of UHECRs in the Galaxy, the anisotropy in all particle spectra is vanishing, except for photons and electrons, for which a small anisotropy in the directions to the Galactic center and opposite one still exists due to synchrotron radiation.
## 2 Propagation through the Galaxy
A numerical codes has been developed recently for calculation of spectra of UHECRs in the case of isotropic and homogeneous source and magnetic field distributions . Our code provides a rather reliable information on possible extragalactic sources of UHECRs. Unfortunately, as was pointed earlier , the method developed in above papers may not be used directly if one is interested in sources within our Galaxy, moreover the galactic magnetic fields may not be neglected even for the extragalactic component of UHECRs in the case of light charged particles (electrons, positrons). Electrons loose their energy rapidly enough due to synchrotron radiation in the galactic magnetic field:
$$\frac{dE}{dt}=\frac{4}{9}\frac{\alpha ^2B^2E^2}{m_e^4},$$
(1)
where $`B10^6G`$ is the galactic magnetic field, and $`E`$ is the UHE electron energy. Then the energy loss length is
$$\left(\frac{1}{E}\frac{dE}{dt}\right)^1=3.8\times \left(\frac{E}{10^{15}eV}\right)^1\left(\frac{B}{10^6G}\right)^2kpc.$$
(2)
Deflection is another important factor when dealing with the propagation problem in a non-anisotropic case. The straight line propagation (SLP) approximation which treats the motion of CR particles in one dimension fails if the effect of the deflection becomes large. The gyroradius of a charged particle with charge $`qe`$ and momentum $`p`$ (energy $`E`$) is given by
$$R_g=\frac{p}{qeB_{}}\frac{E}{qeB_{}}1.1\times \frac{1}{q}\left(\frac{E}{10^{18}\mathrm{eV}}\right)\left(\frac{B_{}}{10^6\mathrm{G}}\right)^1\mathrm{kpc},$$
(3)
where $`B_{}`$ is the field component perpendicular to the particleโs momentum.
However, if one is interested only in the spectra of particles with energies larger than $`10^{19}`$ eV, i.e. particles which weakly (up to 10 degrees) deflect in the galactic magnetic fields $`B10^6`$G on the distances $`10kPc`$, one can use a straight-line propagation approximation. This means that one can use the same kinematic equations as in but written for the differential fluxes instead of concentrations.
$`{\displaystyle \frac{d}{dt}}j_a(E_a,t)=j_a(E_a,t)\alpha _a(E_a,t)+`$ (4)
$`{\displaystyle \underset{c}{}}{\displaystyle \beta _{ca}(E_c,E_a,t)j_c(E_c,t)๐E_c}+{\displaystyle \frac{Q_a(Ee,t)}{4\pi }},`$
where $`j_a(E_a,t)`$ is a differential flux of particles $`a`$.
Integration in Eq. (4) is performed on the chosen straight trajectory within the Galaxy. The initial values of $`j_a(E)`$ may be taken from the previous task (extragalactic component). A some another difference is in a time dependence of external source and magnetic field terms. Their values are determined by current position within the trajectory. Further, for simplicity we will assume spherical symmetry, that is the source concentration and the magnetic field strength are taken to be functions only of distance from the center of the Galaxy $`r`$.
$$B(t)=B^{}(r(t)),Q(t)=Q^{}(r(t)),$$
(5)
As a toy model of galactic magnetic fields we will use the one decreasing exponentially with r:
$$B^{}(r)=B_0exp(r/r_B).$$
(6)
In the Fig.1 we plot point $`E`$, corresponding to the Earth location at a distance $`a`$ from the galactic center $`C`$. Here $`R`$ is an effective Galaxy radius, the distance at which the source concentration and magnetic field strength become equal to extragalactic values. Let us draw an arbitrary trajectory crossing point $`E`$ (horizontal line in the Fig.1), $`\alpha `$ being the angle between the direction to the center and the trajectory and $`t`$ being the length (time) parameter on the trajectory (point $`t=0`$ is the nearest one to the center of Galaxy).
$$r(t)=\sqrt{a^2sin^2\alpha +t^2},$$
(7)
For the flux coming from โcenterโ limits of integration are given by
$$t_{min}=\sqrt{R^2a^2sin^2\alpha },t_{max}=\sqrt{a^2a^2sin^2\alpha };$$
(8)
for the flux from the opposite side
$$t_{min}=\sqrt{a^2a^2sin^2\alpha },t_{max}=\sqrt{R^2a^2sin^2\alpha }.$$
(9)
## 3 Source distributions
As we have already pointed out in Introduction, the absence or presence of sources of UHECR in the Galaxy lead to quite different UHE particle spectra on the Earth. In the first case only electron and photon spectra are slightly anisotropic, while in the latter case an anisotropy is there for all particle spectra.
We considered the same models of source distributions as in . First is the isothermal halo model .
$$n(r)\frac{1}{(r_c^2+r^2)},$$
(10)
where $`r_c`$ is some core size. A model with the distribution (10) may describe the collisionful CDM case. The second model has the distribution of Ref., which corresponds to the opposite limiting case of collisionless CDM
$$n(r)\frac{1}{\sqrt{(r_c^2+r^2)}(R+r)^2}.$$
(11)
We will take always $`r_c=2kpc`$ and $`R=100kpc`$.
Since the distribution of CDM in the Galaxy might contain most probably both collisionful and collisionless components, the results for anisotropy in a realistic case is probably somewhere in between of two extreme cases we considering here.
For the ratio of mean galactic and overall X-particle concentrations we used value
$$\frac{<n_{X_{gal}}>}{<n_X>}=\frac{1}{n_{gal}V_{gal}}=10^5.$$
(12)
## 4 Results
The results of our numerical calculations of propagation of UHECR through the Galaxy for the model of long-living X-particles are presented on Figs.2-6. As an example, we used the following parameters of the model: we took mass of a superheavy particle $`m_X=10^{23}eV=10^{14}GeV`$. We took as the main decay channel the $`Xqq`$ one. We assumed that $`X`$-particles give the main contribution in the Cold Dark Matter (CDM) with $`\mathrm{\Omega }_X=0.3`$ (assuming $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$ and $`H_0=70\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$). We normalized the resulting UHECRs flux to the observed one (AGASA data), and calculated required $`t_X`$ using the fact that fluxes $`j_it_X^1`$. Extragalactic magnetic fields were taken to be equal to $`B_{out}=10^{12}G`$. We used usual QCD-motivated fragmentation functions (see Eq. (52) in ).
We defined anisotropy in spectra of UHECR in the following simple way:
$$A=2\frac{J_{\mathrm{center}}J_{\mathrm{border}}}{J_{\mathrm{center}}+J_{\mathrm{border}}},$$
(13)
where $`J_{\mathrm{center}}`$ and $`J_{\mathrm{border}}`$ are the differential fluxes of UHECR from the direction on the Galactic center and from opposite direction.
We considered two essentially different cases of the $`X`$-particle distribution. In the first case we took uniform distribution of the $`X`$-particles in the Universe, but without $`X`$-particles in our Galaxy. In the second case we considered $`X`$-particle distribution similar to cold dark matter distribution in the galaxies.
On the Fig.2 we plot spectra of UHE protons and photons near the boundary of our Galaxy in case of extragalactic sources of UHECRs. Propagation through the Galaxy mostly leaves the picture at highest energies without changes. It is not so only for electrons, since they are strongly suppressed by galactic magnetic field. Lifetime of $`X`$-particle in this case is $`t_X=8.6\times 10^7t_U`$ ($`t_U`$ is the age of Universe). From Fig.2 one can conclude that all highest energy cosmic rays are photons. This is the general result for all TD models. Moreover, for models with $`m_Xfew\times 10^{22}`$ protons hardly can contribute significantly in the region below cutoff, since an attempt to increase extragalactic magnetic field to fit simultaneously highest energy part of AGASA spectrum by photons and lower energy part by protons will lead to overproducing of photons at EGRET region, due to synchrotron radiation. And so in this case one need to suppose that another kind of UHECR sources exists in order to explain UHECR below cutoff.
The scale of spectra distortions produced during propagating through the Galaxy can be illustrated in terms of the flux asymmetry in directions to and from the Galaxy center Eq.(13). It is presented on the Fig.3. The spectra in any directions are similar for all particles, except for photons: they are a little bit different due to influence of galactic magnetic fields. Namely, some part of photons, which was converted to electrons, was lost due to electrons synchrotron radiation in Galactic magnetic field. Because the path through the Galaxy center is longer then from opposite direction, but both of them much smaller then photon mean free path, there is small asymmetry in photon flux, as we see from Fig.3.
At Fig. 4 we presented typical UHECRs spectra in the case of the galactic origin. As was expected, the UHECR flux in this case is determined by the source distribution rather than by propagation processes. This is not the case only for electrons, because of the extremely small synchrotron energy loss length in magnetic fields of the Galaxy. The $`X`$-particle lifetime is $`t_X=1.5\times 10^{11}t_U`$. We checked that contribution of UHECR from all other galaxies is negligible in this case in compare to the contribution of the Milky Way. The approximate relation for the Galactic versus extragalactic fluxes is given by relation of corresponding $`X`$-particle lifetimes. So, we get $`F_{\mathrm{Galaxy}}/F_{\mathrm{extragalactic}}1.7\times 10^3`$. As we see from Fig. 4, the spectrum of UHECR dominated by photons at all energies. Protons contribute at most up $`10\%`$ to the flux. This is simply due to fragmentation of quarks in $`X`$-particle decay.
In the case of Galactic origin of UHECR the important parameter is anisotropy in particle spectra, defined in Eq. (13). We investigated two essential cases of collisionless Eq.(11) and collisionful Eq.(10) dark matter. The asymmetry in particle spectra for those models is plotted in Fig.5 and Fig.6.
We conclude that anisotropy in spectra of nucleons, photons and neutrino is an important test for top-down models, in which sources are located in galaxies. In the realistic models of source distribution the anisotropy can rise up to 80-90%. The dominant contribution in this case is provided by a flux coming from our Galaxy. The extragalactic input in this case appears to be at least 1000 times smaller. The shape of the spectra at energies higher than few $`\times 10^{19}`$ is determined mostly by source distribution. Deflection in galactic magnetic field is negligible and so the exact configuration of the fields is not important. The flux of UHECR will consist mostly from photons at highest energies.
In the case of the extragalactic UHECRs origin there will remain some anisotropy in photon and electron spectra, however extremely low flux of electrons and low value of photon anisotropy (less than 1%) make the anisotropy hardly detectable. Nevertheless, it is probably worth to notice that in this case more exact models of galactic magnetic fields may change the final result somewhat.
## 5 Acknowledgments
We thank S.L. Dubovsky, P.G. Tinyakov and I.I.Tkachev for fruitful discussion. The work was supported in part by the Russian Fund for Fundamental Research grant 98-02-17493-a and INTAS grant 1A-1065.
|
warning/0006/hep-ph0006249.html
|
ar5iv
|
text
|
# Dimensionally regulated one-loop box scalar integrals with massless internal lines
## 1 Introduction
Scattering processes are one of the most important sources of information on short$``$distance physics and have played a vital role in establishing the fundamental interactions of nature. In testing various aspects of QCD, the scattering processes in which the total number of particles in the initial and final states is $`N5`$ (like 2 $``$ 3, 2 $``$ 4, etc.) are becoming increasingly important.
The techniques for calculating the tree level amplitudes involving a large number of particles in the final state are well established tree . Owing to the well$``$known fact that the LO predictions in perturbative QCD do not have much predictive power, the inclusion of higher$``$order corrections is essential for many reasons. In general, higher- order corrections have a stabilizing effect reducing the dependence of the LO predictions on the renormalization and factorization scales and the renormalization scheme. Therefore, to achieve a complete confrontation between theoretical predictions and experimental data, it is very important to know the size of radiative corrections to the LO predictions.
Obtaining radiative corrections requires evaluation of one$``$loop integrals which arise from a Feynman diagramatic approach. The case of massless internal lines is of special interest, because we often deal with either really massless particles (gluons) or particles whose masses can be neglected in high$``$energy processes (quarks). The main technical difficulty in obtaining the NLO corrections consists in the treatment of the occurring $`N`$$``$point tensor and scalar integrals with massless internal lines. In QCD, tensor integrals appear in which the $`N`$$``$point integral may contain up to $`N`$ powers of the loop momentum in the numerator of the integrand. Since these integrals contain IR divergences, they need to be calculated in an arbitrary number of dimensions and the standard methods of old cannot be directly applied.
Various approaches have been proposed for reducing the dimensionally regulated $`(N5)`$$``$point integrals to a linear combination of $`N`$ and lower$``$point scalar integrals multiplied by tensor structures made from the metric tensor $`g^{\mu \nu }`$ and external momenta davidicev ; dixon ; tarasov ; binoth ; camp .
It has also been shown that the general $`(N5)`$$``$point scalar one$``$loop integral can be recursively represented as a cyclically symmetric combination of $`(N1)`$$``$point integrals, provided the external momenta are kept in four dimensions dixon ; tarasov ; binoth . Consequently, all scalar integrals occurring in the computation of an arbitrary one$``$loop $`(N5)`$$``$point integral with massless internal lines can be reduced to a sum over a set of basic scalar box $`(N=4)`$ integrals with rational coefficients depending on the external momenta and the dimensionality of space$``$time. This set of diagrams includes IR divergent box integrals with massless internal lines but containing 0, 1, 2, and 3 external masses and the IR finite box integral with four external masses.
The IR finite box integral has been evaluated in Ref. hooft , and written in a more compact form in Ref. denner . The results for the IR divergent box integrals with no external masses and with one external mass have been obtained in Refs. fabricius ; papa . All IR divergent box integrals have been considered in Ref. bern , using the partial differential equation technique. However, the results obtained for the integrals containing two and three external masses are strictly correct only in the Euclidean region where all relevant kinematical variables are negative.
Being very nontrivial, impossible to check numerically, and of fundamental importance for one$``$loop calculations in perturbative QCD with massless quarks, it is absolutely essential that these integrals should be evaluated and the results of Ref. bern should be checked using independent techniques.
In this paper we recalculate the IR one$``$loop box scalar integrals in an elegant manner using dimensional regularisation and the Feynman parameter method, and give results in a simple and compact form. A characteristic feature of our calculation is that the causal $`\mathrm{i}ฯต`$ has been systematically kept throughout the calculation, so that the results obtained are valid for arbitrary values of the relevant kinematic variables.
The paper is organized as follows. Section 2 is devoted to introducing the notation and to some preliminary considerations. In Sec. 3, using the Feynman parameter method and the dimensional regularization method, we evaluate the IR divergent one$``$loop box Feynman integrals with massless internal lines but containing 0, 1, 2, and 3 massive external lines, and compare our results with the corresponding ones obtained in Ref. bern . Section 4 is devoted to some concluding remarks. An analytical proof of the equivalence of our results to those obtained in Ref. bern for Euclidean kinematics is given in Appendix A. For the readerโs convenience, in Appendix B we present the closed$``$form expressions for the IR divergent one$``$loop scalar box integrals evaluated in this paper, i.e., with all poles in $`ฯต_{IR}=D/22`$ manifest, and with all functions of the kinematic variables expressed in terms of logarithms and dilogarithms.
## 2 Preliminaries
The massless scalar one$``$loop box integral in $`D`$$``$dimensional space$``$time is given by
$$I_4(p_1,p_2,p_3,p_4)=(\mu ^2)^{2D/2}\frac{\mathrm{d}^Dl}{(2\pi )^D}\frac{1}{A_1A_2A_3A_4},$$
(1)
where $`p_i`$, $`i`$=1,2,3,4 are the external momenta, $`l`$ is the loop momentum, and $`\mu `$ is the usual dimensional regularization scale. As indicated in Fig. 1, all external momenta are taken to be incoming, so that the massless propagators have the form
$`A_1`$ $`=`$ $`l^2+\mathrm{i}ฯต,`$
$`A_2`$ $`=`$ $`(l+p_1)^2+\mathrm{i}ฯต,`$
$`A_3`$ $`=`$ $`(l+p_1+p_2)^2+\mathrm{i}ฯต,`$
$`A_4`$ $`=`$ $`(l+p_1+p_2+p_3)^2+\mathrm{i}ฯต.`$ (2)
Combining the denominators with the help of the Feynman parametrization formula
$`{\displaystyle \frac{1}{A_1A_2A_3A_4}}=`$
$`={\displaystyle _0^1}dx_1dx_2dx_3dx_4{\displaystyle \frac{3!\delta (x_1+x_2+x_3+x_41)}{(x_1A_1+x_2A_2+x_3A_3+x_4A_4)^4}},`$
performing the $`D`$$``$dimensional loop momentum integration using
$$\frac{\mathrm{d}^Dl}{(2\pi )^D}\frac{1}{(l^2M^2+\mathrm{i}ฯต)^4}=\frac{\mathrm{i}}{(4\pi )^{D/2}}\frac{\mathrm{\Gamma }(4D/2)}{3!(M^2\mathrm{i}ฯต)^{4D/2}},$$
(4)
introducing the external โmassesโ
$$p_i^2=m_i^2(i=1,2,3,4),$$
(5)
and the Mandelstam variables
$$s=(p_1+p_2)^2,t=(p_2+p_3)^2,$$
(6)
one readily finds that the scalar integral in (1) can be written in the form
$`I_4(s,t;m_1^2,m_2^2,m_3^2,m_4^2)={\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{\mathrm{\Gamma }(4D/2)}{(4\pi \mu ^2)^{D/22}}}`$ (7)
$`\times {\displaystyle _0^1}\mathrm{d}x_1\mathrm{d}x_2\mathrm{d}x_3\mathrm{d}x_4\delta (x_1+x_2+x_3+x_41)`$
$`\times (x_1x_3sx_2x_4tx_1x_2m_1^2`$
$`x_2x_3m_2^2x_3x_4m_3^2x_1x_4m_4^2\mathrm{i}ฯต)^{D/24}.`$
This is the basic four$``$point โscalarโ parametric integral, serving as a starting point for our further considerations.
Depending on the number of the external massless lines, we distinguish six special cases of this integral. Following the notation of Ref. bern , we denote these integrals by
$`I_4^{4m}`$ $``$ $`I_4(s,t;m_1^2,m_2^2,m_3^2,m_4^2),`$ (8)
$`I_4^{3m}`$ $``$ $`I_4(s,t;0,m_2^2,m_3^2,m_4^2),`$ (9)
$`I_4^{2mh}`$ $``$ $`I_4(s,t;0,0,m_3^2,m_4^2),`$ (10)
$`I_4^{2me}`$ $``$ $`I_4(s,t;0,m_2^2,0,m_4^2),`$ (11)
$`I_4^{1m}`$ $``$ $`I_4(s,t;0,0,0,m_4^2),`$ (12)
$`I_4^{0m}`$ $``$ $`I_4(s,t;0,0,0,0,),`$ (13)
and refer to them as the four$``$mass scalar box integral, the three$``$mass box integral, the โhardโ two$``$mass box integral, the โeasyโ two$``$mass box integral, the one$``$mass box integral, and the massless box integral, respectively.
These six box integrals constitute the fundamental set of integrals, in the sense that an arbitrary one$``$loop $`N(5)`$point integral with massless internal lines can be represented in a unique way as a linear combination of these integrals with the coefficients being rational functions of the relevant kinematic variables and the number of space$``$dimensions $`D`$.
These integrals arise from the Feynman diagrams depicted in Fig. 2, formally corresponding to the scalar massless $`\Phi ^3`$ theory. The thick lines in these diagrams denote the massive (off$``$shell) external lines. As it is seen from Fig. 2, there are two distinct configurations related to the case when two external lines are massless: the adjacent box diagram ($`m_1^2=m_2^2=0`$) and the opposite box diagram ($`m_1^2=m_3^2=0`$). They correspond to the hard and easy two$``$mass box integrals, respectively.
When evaluating the diagrams of Fig. 2, one comes across IR singularities (both collinear and soft). Let us recall the circumstances under which IR singularities appear in a Feynman diagram. When an on$``$shell quark of momentum $`p`$ emits a gluon of momentum $`k`$, then IR singularities appear as a result of vanishing of the quark propagator. If the quark is massless, this can happen when either $`p`$ and $`k`$ are collinear ($`kp`$, collinear singularity) or when the gluon momentum vanishes ($`k0`$, soft singularity). Thus, a Feynman diagram with all particles massless will have soft singularity if it contains an internal gluon line attached to two on$``$shell external quark lines. On the other hand, a diagram will contain collinear singularity if it has an internal gluon line attached to an on$``$shell external quark line. It follows then that a diagram containing soft singularity contains two collinear singularities at the same time, i.e., soft and collinear singularities overlap.
In view of what has been said above, we conclude that the integral $`I_4^{4m}`$ is $`IR`$ finite, and can be calculated in $`D=4`$ space$``$time dimensions, while the rest of integrals contain IR divergence ($`I_4^{3m}`$ and $`I_4^{2me}`$ collinear divergence, $`I_4^{2mh}`$, $`I_4^{1m}`$, and $`I_4^{0m}`$ both collinear and soft divergence), and as such have to be evaluated in $`D=4+2\epsilon _{IR}(\epsilon _{IR}>0)`$ dimensions.
For arbitrary $`D`$, the integrals (9$``$13) cannot be expressed by elementary functions, but we know that the expressions corresponding to these diagrams expanded in powers of $`\epsilon _{IR}`$ are of the generic form
$$\frac{A}{\epsilon _{IR}^2}+\frac{B}{\epsilon _{IR}}+C+๐ช(ฯต_{IR}),$$
with the coefficients $`A,B,C`$ being complex functions of the kinematic invariants. The $`1/\epsilon _{IR}`$ poles express the IR divergence.
As stated in the Introduction, the IR divergent integrals $`I_4^K`$, $`K\{3m,2mh,2me,1m,0m\}`$, defined through Eqs. (7) and (9)$``$(13) , have been evaluated in Ref. bern using the partial differential equation technique. After experimenting with various ways to independently derive the results of Ref. bern , we have found that the Feynman parameter method appears to be the most straightforward and satisfactory approach.
## 3 Calculation and results
Using the Feynman parameter method and the method of dimensional regularization, in this section we evaluate the IR divergent scalar one$``$loop box integrals $`I_4^K`$, $`K\{3m,2mh,2me,1m,0m\}`$. To accomplish that, we start by considering the most complicated of these integrals, namely, the three$``$mass box integral $`I_4^{3m}`$. We show that, paying due attention to the fact that the limit of taking a mass to zero does not necessarily commute with the $`\epsilon _{IR}`$ expansion of dimensional regularization, the results obtained at the intermediate steps of the calculation of the integral $`I_4^{3m}`$ can be used to obtain the results for the rest of the above integrals. A characteristic feature of our calculation is that we keep the causal $`\mathrm{i}ฯต`$ systematically through the calculation, so that the results we obtain are valid for arbitrary values of the relevant kinematic variables.
Let us then start with the three$``$mass box integral $`I_4^{3m}`$. By setting $`m_1^2=0`$ and $`D=4+2\epsilon _{IR}(\epsilon _{IR}>0)`$ in Eq. (7) and eliminating the $`\delta `$function by performing the $`x_4`$ integration the integral $`I_4^{3m}`$ becomes
$`I_4^{3m}`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{\mathrm{\Gamma }(2\epsilon _{IR})}{(4\pi \mu ^2)^{\epsilon _{IR}}}}{\displaystyle _0^1}dx_1{\displaystyle _0^{1x_1}}dx_2{\displaystyle _0^{1x_1x_2}}dx_3`$ (14)
$`\times [x_1x_3sx_2x_3m_2^2(1x_1x_2x_3)`$
$`\times (x_1m_4^2+x_2t+x_3m_3^2)\mathrm{i}ฯต]^{\epsilon _{IR}2}.`$
It is a well$``$known fact that the appropriate choice of Feynman parameters is in practice a critical ingredient in enabling one to evaluate a complicated Feynman integral analytically. There does not appear to be any simple formula for choosing an optimal set of Feynman parameters for a given diagram, as there is generally an enormous set of possibilities.
To proceed with the evaluation of the integral $`I_4^{3m}`$, the most suitable set of Feynman parameters turns out to be given by karplus
$`x_1`$ $`=`$ $`(1x)(1y),`$
$`x_2`$ $`=`$ $`x(1y),`$
$`x_3`$ $`=`$ $`yz.`$ (15)
The Jacobian corresponding to this transformation of the integration variables is $`y(1y)`$. Written in terms of the new variables, the integral (14) takes the form
$`I_4^{3m}={\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{\mathrm{\Gamma }(2\epsilon _{IR})}{(4\pi \mu ^2)^{\epsilon _{IR}}}}{\displaystyle _0^1}\mathrm{d}x\mathrm{d}y\mathrm{d}zy(1y)\{y(1y)`$ (16)
$`\times \left[(1x)zs+(1x)(1z)m_4^2+xzm_2^2+x(1z)t\right]`$
$`z(1z)y^2m_3^2\mathrm{i}ฯต\}^{\epsilon _{IR}2}.`$
As it is seen from Eq. (16), the integration over $`x`$ is elementary and is readily performed. The resulting expression is
$`I_4^{3m}`$ $`=`$ $`{\displaystyle \frac{\kappa }{2}}{\displaystyle _0^1}dydz{\displaystyle \frac{1}{z(sm_2^2)+(1z)(m_4^2t)}}`$ (17)
$`\times \{[y(1y)(zs+(1z)m_4^2)`$
$`z(1z)y^2m_3^2\mathrm{i}ฯต]^{\epsilon _{IR}1}`$
$`[y(1y)(zm_2^2+(1z)t)`$
$`z(1z)y^2m_3^2\mathrm{i}ฯต]^{\epsilon _{IR}1}\},`$
where we have introduced the abbreviation
$$\kappa =\frac{\mathrm{i}}{(4\pi )^2}\frac{2\mathrm{\Gamma }(1\epsilon _{IR})}{(4\pi \mu ^2)^{\epsilon _{IR}}}$$
(18)
Next, by pulling out the factor $`y^{\epsilon _{IR}1}`$ from both terms in the curly brackets (which is legitimate since $`y`$ is positive in the integration region and the sign of $`ฯต(>0)`$ does not change), and making use of the following relation:
$`(a+b\mathrm{i}ฯต)^{\epsilon _{IR}1}`$ $`=`$ $`\left({\displaystyle \frac{a}{b\mathrm{i}ฯต}}+1\right)^{\epsilon _{IR}1}(b\mathrm{i}ฯต)^{\epsilon _{IR}1}`$ (19)
$`(a,b๐),`$
(which is not self$``$evident owing to the fact that $`\epsilon _{IR}`$ is not an integer) leads to the integral of the form
$`I_4^{3m}={\displaystyle \frac{\kappa }{2}}{\displaystyle _0^1}dz{\displaystyle \frac{1}{z(sm_2^2)+(1z)(m_4^2t)}}`$ (20)
$`\times \{(zs(1z)m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}1}{\displaystyle _0^1}\mathrm{d}yy^{\epsilon _{IR}1}`$
$`\times \left[1y\left(1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zs+(1z)m_4^2+\mathrm{i}ฯต}}\right)\right]^{\epsilon _{IR}1}`$
$`\left(zm_2^2(1z)t\mathrm{i}ฯต\right)^{\epsilon _{IR}1}{\displaystyle _0^1}dyy^{\epsilon _{IR}1}`$
$`\times [1y(1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zm_2^2+(1z)t+\mathrm{i}ฯต}})]^{\epsilon _{IR}1}\}.`$
By noticing that the integral over $`y`$ stands for the Euler integral representation of the hypergeometric function
$`{}_{2}{}^{}F_{1}^{}(a,b;c;z)`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(c)}{\mathrm{\Gamma }(b)\mathrm{\Gamma }(cb)}}{\displaystyle _0^1}dt{\displaystyle \frac{t^{b1}(1t)^{cb1}}{(1tz)^a}},`$ (21)
$`\text{Re}c>\text{Re}b>0;\text{arg}(1z)<\pi ,`$
we obtain the result
$`I_4^{3m}={\displaystyle \frac{\kappa }{2\epsilon _{IR}}}{\displaystyle _0^1}dz{\displaystyle \frac{1}{z(sm_2^2)+(1z)(m_4^2t)}}`$
$`\times [(zs(1z)m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}1}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zs+(1z)m_4^2+\mathrm{i}ฯต}})`$
$`\left(zm_2^2(1z)t\mathrm{i}ฯต\right)^{\epsilon _{IR}1}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zm_2^2+(1z)t+\mathrm{i}ฯต}})].`$
Next, using the identity (partial fraction decomposition)
$$\frac{1}{az+b}\frac{1}{cz+d}=\frac{1}{adbc}\left(\frac{a}{az+b}\frac{c}{cz+d}\right),$$
(23)
and performing a few simple rearrangements, we find that the integral under consideration can be written in the following form:
$$I_4^{3m}=\frac{\kappa }{stm_2^2m_4^2}(P^{3m}+Q^{3m}),$$
(24)
where
$`P^{3m}={\displaystyle \frac{1}{2\epsilon _{IR}}}[(m_2^2t){\displaystyle _0^1}\mathrm{d}z[zm_2^2(1z)t\mathrm{i}ฯต]^{\epsilon _{IR}1}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zm_2^2+(1z)t+\mathrm{i}ฯต}})`$
$`+(m_4^2s){\displaystyle _0^1}dz[zs(1z)m_4^2\mathrm{i}ฯต]^{\epsilon _{IR}1}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zs+(1z)m_4^2+\mathrm{i}ฯต}})],`$
and
$`Q^{3m}={\displaystyle \frac{1}{2\epsilon _{IR}}}{\displaystyle _0^1}dz{\displaystyle \frac{1}{zz_0}}`$
$`\times [(zm_2^2(1z)t\mathrm{i}ฯต)^{\epsilon _{IR}}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zm_2^2+(1z)t+\mathrm{i}ฯต}})`$
$`(zs(1z)m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{zs+(1z)m_4^2+\mathrm{i}ฯต}})],`$
with
$$z_0=\frac{tm_4^2}{s+tm_2^2m_4^2}.$$
(27)
In view of (24), it is clear that the integrals $`I_4^K`$ can be written in the form
$`I_4^K`$ $`=`$ $`\kappa g^K(P^K+Q^K),`$ (28)
$`K\{3m,2mh,2me,1m,0m\},`$
with
$$g^{3m}=g^{2me}=\frac{1}{stm_2^2m_4^2},g^{2mh}=g^{1m}=g^{0m}=\frac{1}{st}$$
(29)
Not being able to perform the remaining integrations in Eqs. (LABEL:eq:f25) and (LABEL:eq:f27) analytically, we proceed by expanding the integrands in power series in $`\epsilon _{IR}`$ and by term by term integration. We shall see below that all divergences of the integral $`I_4^K`$ are contained in $`P^K`$, while $`Q^K`$ is completely finite.
It is clear from Eqs. (18) and (24) that to obtain the values of the integrals $`I_4^K`$, $`K\{3m,2mh,2me,1m,0m\}`$ to order $`๐ช(\epsilon _{IR}^0)`$, the evaluation of the integrals $`P^K`$ and $`Q^K`$ should be made to the same order.
It turns out, however, that the number of integrals that really need to be evaluted reduces to just four, namely, $`P^{3m}`$, given by (LABEL:eq:f27), the integrals $`P^{2mh}`$ and $`P^{2me}`$, obtained by setting $`m_2^2=0`$, and $`m_3^2=0`$ in (LABEL:eq:f27), respectively, and the integral $`Q^{3m}`$ given by (LABEL:eq:f25). The values of all the other integrals appearing in Eqs. (28) can be derived by taking appropriate zero$``$mass limits.
### 3.1 Calculation of the integrals $`P^{3m}`$, $`P^{2mh}`$, and $`P^{2me}`$
Let us start by considering the integral $`P^{3m}`$ given by Eq. (LABEL:eq:f27). Making the change $`z1z`$ in the second term on the right$``$hand side in (LABEL:eq:f27), one finds that the first and second terms are related to each other by the $`ts`$ and $`m_2^2m_4^2`$ interchanges. Therefore, the expression for $`P^{3m}`$ takes the form
$$P^{3m}=\frac{1}{2}\left[R(t,m_2^2,m_3^2)+R(s,m_4^2,m_3^2)\right],$$
(30)
where
$`R(\alpha ,\beta ,m_3^2)=`$
$`={\displaystyle \frac{1}{\epsilon _{IR}}}{\displaystyle _0^1}dz(\beta \alpha )(z\beta (1z)\alpha \mathrm{i}ฯต)^{\epsilon _{IR}1}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{z\beta +(1z)\alpha +\mathrm{i}ฯต}}).`$
In order to evaluate this integral, we first apply the linear transformation formula for the hypergeometric functions:
$`{}_{2}{}^{}F_{1}^{}(a,b;c;z)`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(c)\mathrm{\Gamma }(cab)}{\mathrm{\Gamma }(ca)\mathrm{\Gamma }(cb)}}`$ (32)
$`\times {}_{2}{}^{}F_{1}^{}(a,b;1+a+bc;1z)`$
$`+{\displaystyle \frac{\mathrm{\Gamma }(c)\mathrm{\Gamma }(a+bc)}{\mathrm{\Gamma }(a)\mathrm{\Gamma }(b)}}(1z)^{cab}`$
$`\times {}_{2}{}^{}F_{1}^{}(ca,cb;1+cab;1z)`$
$`cab\pm n,\text{arg}(1z)<\pi .`$
With Eq. (32) taken into account, (LABEL:eq:f46) becomes
$`R(\alpha ,\beta ,m_3^2)=`$ (33)
$`={\displaystyle \frac{1}{\epsilon _{IR}}}{\displaystyle _0^1}dz(\beta \alpha )(z\beta (1z)\alpha \mathrm{i}ฯต)^{\epsilon _{IR}1}`$
$`\times [{\displaystyle \frac{\mathrm{\Gamma }(1+\epsilon _{IR})\mathrm{\Gamma }(\epsilon _{IR})}{\mathrm{\Gamma }(2\epsilon _{IR})}}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1\epsilon _{IR};{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{z\beta +(1z)\alpha +\mathrm{i}ฯต}})`$
$`\left({\displaystyle \frac{z(1z)m_3^2\mathrm{i}ฯต}{z\beta (1z)\alpha \mathrm{i}ฯต}}\right)^{\epsilon _{IR}}`$
$`\times {}_{2}{}^{}F_{1}^{}(2\epsilon _{IR},1;1+\epsilon _{IR};{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{z\beta +(1z)\alpha +\mathrm{i}ฯต}})].`$
Assuming, with no loss of generality, that
$$m_3^2<\alpha ,\beta \alpha /\beta >0,$$
(34)
from which it follows
$$\left|\frac{z(1z)m_3^2+\mathrm{i}ฯต}{z\beta +(1z)\alpha +\mathrm{i}ฯต}\right|<1z[0,1],$$
(35)
and making use of the series representation of the hypergeometric function
$${}_{2}{}^{}F_{1}^{}(a,b;c;z)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{\mathrm{\Gamma }(a+n)\mathrm{\Gamma }(b+n)\mathrm{\Gamma }(c)}{\mathrm{\Gamma }(a)\mathrm{\Gamma }(b)\mathrm{\Gamma }(c+n)}\frac{z^n}{n!},z<1,$$
(36)
leads to the result
$`R(\alpha ,\beta ,m_3^2)={\displaystyle \frac{\mathrm{\Gamma }(\epsilon _{IR})}{\mathrm{\Gamma }(2\epsilon _{IR})}}\left(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right){\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right)^n`$
$`\times \{(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \frac{\mathrm{\Gamma }(n+\epsilon _{IR})}{\mathrm{\Gamma }(1+n)}}`$
$`\times {\displaystyle _0^1}\mathrm{d}z\left[z(1z)\right]^n[1z(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})]^{(1+n\epsilon _{IR})}`$
$`(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \frac{\mathrm{\Gamma }(n+2\epsilon _{IR})}{\mathrm{\Gamma }(1+n+\epsilon _{IR})}}`$
$`\times {\displaystyle _0^1}\mathrm{d}z\left[z(1z)\right]^{n+\epsilon _{IR}}[1z(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})]^{(1+n)}\}.`$
Taking into account the integral representation of the hypergeometric function given by Eq. (21) yields
$`R(\alpha ,\beta ,m_3^2)={\displaystyle \frac{\mathrm{\Gamma }(\epsilon _{IR})}{\mathrm{\Gamma }(2\epsilon _{IR})}}\left(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right){\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right)^n`$
$`\times [(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \frac{\mathrm{\Gamma }(n+\epsilon _{IR})\mathrm{\Gamma }(1+n)}{\mathrm{\Gamma }(2+2n)}}`$
$`\times {}_{2}{}^{}F_{1}^{}(1+n\epsilon _{IR},1+n;2+2n;1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})`$
$`(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \frac{\mathrm{\Gamma }(n+2\epsilon _{IR})\mathrm{\Gamma }(1+n+\epsilon _{IR})}{\mathrm{\Gamma }(2+2n+2\epsilon _{IR})}}`$
$`\times {}_{2}{}^{}F_{1}^{}(1+n,1+n+\epsilon _{IR};2+2n+2\epsilon _{IR};1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})].`$
Although obtained under the restriction expressed by Eq. (34), the above expression for $`R(\alpha ,\beta ,m_3^2)`$ can be analytically continued for an arbitrary value of $`\beta `$. The series in (LABEL:eq:f53) is convergent because it converges for $`\beta =0`$ (see below) and the hypergeometric functions reach the maximum at the same point. To retain the convergence of the series, we still assume that $`m_3^2<\alpha `$.
By inspecting the sum on the right$``$hand side of (LABEL:eq:f53), we find that, if $`\beta 0`$, all terms with the exception of the first one ($`n=0`$) are of higher order in $`\epsilon _{IR}`$ and can therefore be omitted. Upon replacing the whole sum by the first term, and performing a few simple rearrangements, we find
$`R(\alpha ,\beta 0,m_3^2)={\displaystyle \frac{\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(2\epsilon _{IR})}}(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})[(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}`$
$`\times {}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},1;2;1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})`$
$`(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \frac{1}{2(1+2\epsilon _{IR})}}`$
$`\times {}_{2}{}^{}F_{1}^{}(1,1+\epsilon _{IR};2+2\epsilon _{IR};1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})]+๐ช(\epsilon _{IR}).`$
On the basis of formula (21), one can easily show that the first hypergeometric function on the right$``$hand side of the above expression can be written in the form
$`{}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},1;2;1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})=`$
$`={\displaystyle \frac{1}{\epsilon _{IR}}}\left(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right)^1{\displaystyle \frac{(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}(\beta \mathrm{i}ฯต)^{\epsilon _{IR}}}{(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}}}.`$
Now, inserting (3.1) into (LABEL:eq:f54) leads to the expression
$`R(\alpha ,\beta ,m_3^2)={\displaystyle \frac{2\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}}[(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}(\beta \mathrm{i}ฯต)^{\epsilon _{IR}}`$
$`{\displaystyle \frac{(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}}{2}}{\displaystyle \frac{\epsilon _{IR}}{1+2\epsilon _{IR}}}\left(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right)`$
$`\times {}_{2}{}^{}F_{1}^{}(1,1+\epsilon _{IR};2+2\epsilon _{IR};1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})]+๐ช(\epsilon _{IR}),`$
which is valid for arbitrary values of $`\alpha `$ and $`m_3^2`$, and for $`\beta 0`$.
On the other hand, if $`\beta =0`$, the first and second hypergeometric functions appearing on the right$``$hand side of Eq. (LABEL:eq:f53) reduce to
$$\frac{\mathrm{\Gamma }(2+2n)\mathrm{\Gamma }(\epsilon _{IR})}{\mathrm{\Gamma }(1+n)\mathrm{\Gamma }(1+n+\epsilon _{IR})}$$
and
$$\frac{\mathrm{\Gamma }(2+2n+2\epsilon _{IR})\mathrm{\Gamma }(\epsilon _{IR})}{\mathrm{\Gamma }(1+n+2\epsilon _{IR})\mathrm{\Gamma }(1+n+\epsilon _{IR})},$$
respectively. This implies that when $`\beta =0`$, all terms in the sum are individually divergent and of the same order in $`\epsilon _{IR}`$ and, as such, all have to be taken into account, i. e., the summation has to be performed explicitly. Therefore, as far as the expression for $`R(\alpha ,\beta ,m_3^2)`$ given by Eq. (LABEL:eq:f53) is concerned, the cases $`\beta =0`$ and $`\beta 0`$ ought to be considered separately.
By setting $`\beta =0`$, Eq. (LABEL:eq:f53) takes the form
$`R(\alpha ,0,m_3^2)=`$ (42)
$`={\displaystyle \frac{\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(2\epsilon _{IR})}}[(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n+\epsilon _{IR}}}\left({\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right)^n`$
$`(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n+2\epsilon _{IR}}}\left({\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}}\right)^n].`$
For the purpose of performing the summations in (42), we note that the series representation of the hypergeometric function given by (36) particularized for $`a=1,c=1+b`$ leads to the formula
$${}_{2}{}^{}F_{1}^{}(1,b;1+b;z)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{b}{n+b}z^nz<1,$$
(43)
which, when taken into account in (42), leads to the result
$`R(\alpha ,0,m_3^2)={\displaystyle \frac{2\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}}`$
$`\times [(\alpha \mathrm{i}ฯต)_2^{\epsilon _{IR}}F_1(1,\epsilon _{IR};1+\epsilon _{IR};{\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})`$
$`{\displaystyle \frac{1}{2}}(m_3^2\mathrm{i}ฯต)_2^{\epsilon _{IR}}F_1(1,2\epsilon _{IR};1+2\epsilon _{IR};{\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})].`$
This expression for $`R(\alpha ,0,m_3^2)`$ is valid for arbitrary values of $`\alpha `$ and $`m_3^2`$.
To establish contact with the results obtained in Ref. bern , instead of expanding everything in the expressions (LABEL:eq:f56) and (LABEL:eq:f59) to the required order in $`\epsilon _{IR}`$, let us, for the time being, expand only the occurring hypergeometric functions. The relevant expansions are
$`z_2F_1(1,1+\delta ;2+2\delta ;z)=`$ (45)
$`={\displaystyle \frac{1+2\delta }{\delta }}\left[1(1z)^\delta 2\delta ^2\text{Li}_2(z)+๐ช(\delta ^3)\right]`$
and
$${}_{2}{}^{}F_{1}^{}(1,\delta ;1+\delta ;z)=1\delta \mathrm{ln}(1z)\delta ^2\text{Li}_2(z)+๐ช(\delta ^3),$$
(46)
where $`\text{Li}_2(z)`$ stands for the Euler dilogarithm hill defined as
$$\mathrm{Li}_2(z)=_0^1\frac{dt}{t}\mathrm{ln}(1tz).$$
(47)
Taking (45) into account in (LABEL:eq:f56), we find that
$`R(\alpha ,\beta ,m_3^2)={\displaystyle \frac{2\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}}[(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}(\beta \mathrm{i}ฯต)^{\epsilon _{IR}}`$ (48)
$`{\displaystyle \frac{1}{2}}(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{(\beta \mathrm{i}ฯต)^{\epsilon _{IR}}(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}}{(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}}}`$
$`+\epsilon _{IR}^2\text{Li}_2(1{\displaystyle \frac{\beta +\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})]+๐ช(\epsilon _{IR}).`$
Next, using the expansion (46), the expression (LABEL:eq:f59) becomes
$`R(\alpha ,0,m_3^2)=`$ (49)
$`={\displaystyle \frac{2\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}}\{(\alpha \mathrm{i}ฯต)^{\epsilon _{IR}}{\displaystyle \frac{1}{2}}(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}`$
$`+\epsilon _{IR}^2[\text{Li}_2(1{\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{\alpha +\mathrm{i}ฯต}})+{\displaystyle \frac{\pi ^2}{6}}]\}+๐ช(\epsilon _{IR}).`$
In arriving at (49), the relation
$$\text{Li}_2(1z)=\text{Li}_2(z)\mathrm{ln}(z)\mathrm{ln}(1z)+\frac{\pi ^2}{6},$$
(50)
has been employed. On the basis of Eqs. (48) and (30) we find that the integral $`P^{3m}`$ is given by
$`P^{3m}={\displaystyle \frac{\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}}\{(s\mathrm{i}ฯต)^{\epsilon _{IR}}+(t\mathrm{i}ฯต)^{\epsilon _{IR}}`$ (51)
$`(m_2^2\mathrm{i}ฯต)^{\epsilon _{IR}}(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}(m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}}`$
$`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{(m_2^2\mathrm{i}ฯต)^{\epsilon _{IR}}(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}}{(t\mathrm{i}ฯต)^{\epsilon _{IR}}}}`$
$`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}(m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}}}{(s\mathrm{i}ฯต)^{\epsilon _{IR}}}}`$
$`+\epsilon _{IR}^2[\text{Li}_2(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})+\text{Li}_2(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}})]\}`$
$`+๐ช(\epsilon _{IR}).`$
Similarly, combining Eqs. (48), (49), and (30), we obtain
$`P^{2mh}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}}\{(s\mathrm{i}ฯต)^{\epsilon _{IR}}+(t\mathrm{i}ฯต)^{\epsilon _{IR}}`$ (52)
$`(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}(m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}}`$
$`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{(m_3^2\mathrm{i}ฯต)^{\epsilon _{IR}}(m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}}}{(s\mathrm{i}ฯต)^{\epsilon _{IR}}}}`$
$`+\epsilon _{IR}^2[{\displaystyle \frac{\pi ^2}{6}}\text{Li}_2(1{\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})`$
$`+\text{Li}_2(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}})]\}+๐ช(\epsilon _{IR}).`$
We now turn to evaluate the integral $`P^{2me}`$. To this end, we set $`m_3^2=0`$ in Eq. (LABEL:eq:f27). As a result, both hypergeometric functions appearing in (LABEL:eq:f27) reduce to
$${}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1)=\frac{\mathrm{\Gamma }(1+\epsilon _{IR})\mathrm{\Gamma }(\epsilon _{IR})}{\mathrm{\Gamma }(2\epsilon _{IR})},$$
(53)
making it possible to perform the remaining integration analytically. The exact result for the integral $`P^{2me}`$ is
$`P^{2me}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }^2(\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}}[(s\mathrm{i}ฯต)^{\epsilon _{IR}}+(t\mathrm{i}ฯต)^{\epsilon _{IR}}`$ (54)
$`(m_2^2\mathrm{i}ฯต)^{\epsilon _{IR}}(m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}}].`$
It is valid for arbitrary values of $`m_2^2`$ and $`m_4^2`$. Consequently, the expressions for the integrals $`P^{1m}`$ and $`P^{0m}`$ are obtained by setting $`m_2^2=0`$ and $`m_2^2=m_4^2=0`$, respectively, in (54).
A remark concerning the issue of the zero$``$mass limits of the massive integrals $`P^K`$ is in order. By looking at the expressions for the integrals $`P^K`$ given above, one observes that the limits $`P^{2mh}P^{1m}`$, $`P^{3m}P^{2me}`$, and $`P^{3m}P^{2mh}`$ are not smooth. On the other hand, the limits $`P^{2me}P^{1m}`$ and $`P^{1m}P^{0m}`$ are smooth. In general, there is no reason for the zero$``$ mass limits to be smooth. Namely, the limit of taking a mass to zero does not necessarily commute with the $`1/\epsilon _{IR}`$ expansion of the dimensional regularization, which has been truncated at $`๐ช(\epsilon _{IR}^0)`$. Note, however, that if we were able to evaluate the integral $`P^{3m}`$ in (LABEL:eq:f27) analytically for general $`\epsilon _{IR}`$, the result thus obtained would suffice to obtain the results for the other integrals $`P^{2mh},P^{2me},P^{1m}`$, and $`P^{0m}`$ by simply setting $`m_2^2=0`$, $`m_3^2=0`$, $`m_2^2=m_3^2=0`$, and $`m_2^2=m_3^2=m_4^2=0`$, respectively, and then expanding these results to the required order in $`\epsilon _{IR}`$.
### 3.2 Calculation of the integral $`Q^{3m}`$
As it is seen from (LABEL:eq:f25), the integral $`Q^{3m}`$ is given in terms of two hypergeometric functions, both of which can be conveniently written in the form
$`{}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{z\alpha +(1z)\beta +\mathrm{i}ฯต}})`$
$`(\alpha ,\beta )\{(m_2^2,t),(s,m_4^2)\}.`$ (55)
Making use of the transformation formula
$${}_{2}{}^{}F_{1}^{}(a,b;c;z)=(1z)_2^{cab}F_1(ca,cb;c;z),$$
(56)
the symmetry of $`{}_{2}{}^{}F_{1}^{}`$ with respect to the arguments $`a`$ and $`b`$, i.e.,
$${}_{2}{}^{}F_{1}^{}(a,b;c;z)=_2F_1(b,a;c;z),$$
(57)
and the integral representation of the hypergeometric function (21), we can write the hypergeometric function given by (55) in the form
$`{}_{2}{}^{}F_{1}^{}(1\epsilon _{IR},\epsilon _{IR};1+\epsilon _{IR};1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{z\alpha +(1z)\beta +\mathrm{i}ฯต}})=`$
$`=\left({\displaystyle \frac{z(1z)m_3^2\mathrm{i}ฯต}{z\alpha (1z)\beta \mathrm{i}ฯต}}\right)^{\epsilon _{IR}}{\displaystyle \frac{\mathrm{\Gamma }(1+\epsilon _{IR})}{\mathrm{\Gamma }(1\epsilon _{IR})\mathrm{\Gamma }(2\epsilon _{IR})}}`$
$`\times {\displaystyle _0^1}\mathrm{d}yy^{2\epsilon _{IR}1}(1y)^{\epsilon _{IR}}`$
$`\times \left[1y\left(1{\displaystyle \frac{z(1z)m_3^2+\mathrm{i}ฯต}{z\alpha +(1z)\beta +\mathrm{i}ฯต}}\right)\right]^1.`$ (58)
Upon substituting (58) into (LABEL:eq:f25), and utilizing the identity (23) once more, we find that
$`Q^{3m}={\displaystyle \frac{\mathrm{\Gamma }(\epsilon _{IR})}{2\mathrm{\Gamma }(1\epsilon _{IR})\mathrm{\Gamma }(2\epsilon _{IR})}}{\displaystyle _0^1}dz{\displaystyle \frac{1}{zz_0}}`$ (59)
$`\times [z(1z)m_3^2\mathrm{i}ฯต]^{\epsilon _{IR}}{\displaystyle _0^1}dyy^{2\epsilon _{IR}}(1y)^{\epsilon _{IR}}`$
$`\times \{[z(1z)m_3^2+zs+(1z)m_4^2]`$
$`\times [zs(1z)m_4^2\mathrm{i}ฯต`$
$`+y(z(1z)m_3^2+zs+(1z)m_4^2)]^1`$
$`\left[z(1z)m_3^2+zm_2^2+(1z)t\right]`$
$`\times [zm_2^2(1z)t\mathrm{i}ฯต`$
$`+y(z(1z)m_3^2+zm_2^2+(1z)t)]^1\}.`$
Since we are interested in obtaining the value of $`Q^{3m}`$ to $`๐ช(\epsilon _{IR}^0)`$, the fact that the expansion of the prefactor in the above expression is of the form $`1+๐ช(\epsilon _{IR})`$, allows us to set $`\epsilon _{IR}=0`$ in the integrand in (59). As a result, the expression for $`Q^{3m}`$ reduces to
$`Q^{3m}={\displaystyle _0^1}dz{\displaystyle \frac{1}{zz_0}}{\displaystyle _0^1}dy`$
$`\times \{[z(1z)m_3^2+zs+(1z)m_4^2]`$
$`\times [zs(1z)m_4^2\mathrm{i}ฯต`$
$`+y(z(1z)m_3^2+zs+(1z)m_4^2)]^1`$
$`\left[z(1z)m_3^2+zm_2^2+(1z)t\right]`$
$`\times [zm_2^2(1z)t\mathrm{i}ฯต`$
$`+y(z(1z)m_3^2+zm_2^2+(1z)t)]^1\}+๐ช(\epsilon _{IR}).`$
Performing the $`y`$ integration, we find
$$Q^{3m}=_0^1\frac{\mathrm{d}z}{zz_0}\mathrm{ln}\left(\frac{z(tm_2^2)t\mathrm{i}ฯต}{z(m_4^2s)m_4^2\mathrm{i}ฯต}\right)+๐ช(\epsilon _{IR})$$
(61)
To carry out the remaining integration, it is important to note that the residue of the integrand at the pole $`z_0`$ is zero, and the logarithm does not cross the cut. This fact allows us to make a few simple transformations of the integrand.
Thus, upon the substitution $`zz+z_0`$, the decomposition
$$_0^1๐z_0^{z_0}๐z+_0^{1z_0}๐z,$$
followed by a change of the variable $`zzz_0`$ in the first, and $`zz(1z_0)`$ in the second term, the integral $`Q^{3m}`$ can be written down as
$`Q^{3m}`$ $`=`$ $`{\displaystyle _0^1}{\displaystyle \frac{\mathrm{d}z}{z}}\{\mathrm{ln}[1z(1(m_2^2+\mathrm{i}ฯต)f^{3m})]`$ (62)
$`+\mathrm{ln}\left[1z\left(1(m_4^2+\mathrm{i}ฯต)f^{3m}\right)\right]`$
$`\mathrm{ln}\left[1z\left(1(s+\mathrm{i}ฯต)f^{3m}\right)\right]`$
$`\mathrm{ln}[1z(1(t+\mathrm{i}ฯต)f^{3m})]\}+๐ช(\epsilon _{IR}),`$
where we have introduced the abbreviation
$$f^{3m}=\frac{s+tm_2^2m_4^2}{stm_2^2m_4^2}.$$
(63)
Expressed in terms of the Euler dilogarithm, the final result for the integral $`Q^{3m}`$ takes the form
$`Q^{3m}=\text{Li}_2\left[1(s+\mathrm{i}ฯต)f^{3m}\right]+\text{Li}_2\left[1(t+\mathrm{i}ฯต)f^{3m}\right]`$
$`\text{Li}_2[1(m_2^2+\mathrm{i}ฯต)f^{3m}]\text{Li}_2[1(m_4^2+\mathrm{i}ฯต)f^{3m}]+๐ช(\epsilon _{IR}).`$
The expression for $`Q^{3m}`$ does not depend on $`m_3^2`$, and is valid for arbitrary values of $`m_2^2`$ and $`m_4^2`$. A consequence of this is that the expression for the integral $`Q^{2mh}`$ to the same order in $`\epsilon _{IR}`$ can simply be obtained by setting $`m_2^2=0`$ in (3.2). Note, however, that, strictly speaking, the expressions for $`Q^{2me}`$, $`Q^{1m}`$, and $`Q^{0m}`$ cannot be obtained by taking appropriate zero$``$mass limits of the same expression. Namely, the expression (59) has been derived assuming that $`m_3^20`$. Therefore, the $`\epsilon _{IR}`$ expansion of (59) is not justified in the $`m_3^20`$ limit. In order to obtain the integrals $`Q^{2me}`$, $`Q^{1m}`$, and $`Q^{0m}`$, we proceed as follows. We return to Eq. (LABEL:eq:f25) and set $`m_3^2=0`$. As a result, Eq. (LABEL:eq:f25) reduces to
$`Q^{2me}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }^2(\epsilon _{IR})}{2\mathrm{\Gamma }(2\epsilon _{IR})}}{\displaystyle _0^1}\mathrm{d}z{\displaystyle \frac{1}{zz_0}}[(zm_2^2(1z)t\mathrm{i}ฯต)^{\epsilon _{IR}}`$ (65)
$`(zs(1z)m_4^2\mathrm{i}ฯต)^{\epsilon _{IR}}].`$
Expanding this expression into power series in $`\epsilon _{IR}`$
$$Q^{2me}=_0^1\frac{\mathrm{d}z}{zz_0}\mathrm{ln}(\frac{z(tm_2^2)t\mathrm{i}ฯต}{z(m_4^2s)m_4^2\mathrm{i}ฯต})+๐ช(\epsilon _{IR}),$$
(66)
and comparing it with the expansion of $`Q^{3m}`$ given by Eq. (61), we find that the expansions for $`Q^{3m}`$ and $`Q^{2me}`$ coincide to the required order in $`\epsilon _{IR}`$. This being the case, all other intgerals $`Q^K`$ can be derived from the integral $`Q^{3m}`$ given by Eq. (3.2) by taking appropriate zero$``$mass limits.
### 3.3 Results
Having obtained, in the preceding subsections, the closed form expressions for the integrals $`P^K(s,t;\{m_i^2\})`$ and $`Q^K(s,t;\{m_i^2\})`$ to order $`๐ช`$$`(\epsilon _{IR}^0)`$, we are now in a position to write down explicit expressions for the integrals $`I_4^K(s,t;\{m_i^2\})`$.
Before proceeding, it is convenient to introduce the functions
$`f^{3m}=f^{2me}`$ $`=`$ $`{\displaystyle \frac{s+tm_2^2m_4^2}{stm_2^2m_4^2}},`$
$`f^{2mh}=f^{1m}`$ $`=`$ $`{\displaystyle \frac{s+tm_4^2}{st}},`$
$`f^{0m}`$ $`=`$ $`{\displaystyle \frac{s+t}{st}},`$ (67)
and the notation
$$r_\mathrm{\Gamma }=\frac{\mathrm{\Gamma }(1\epsilon _{IR})\mathrm{\Gamma }^2(1+\epsilon _{IR})}{\mathrm{\Gamma }(1+2\epsilon _{IR})}$$
(68)
for the $`\mathrm{\Gamma }`$ function prefactor.
Now, on the basis of Eqs. (28), (3.2), (51), (52), and (54), we find that
the three$``$mass scalar box integral is
$`I_4^{3m}(s,t;m_2^2,m_3^2,m_4^2)={\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{r_\mathrm{\Gamma }}{stm_2^2m_4^2}}`$ (69)
$`\times \{{\displaystyle \frac{2}{\epsilon _{IR}^2}}[\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}+\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_2^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$
$`\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_4^2iฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}]`$
$`+{\displaystyle \frac{1}{\epsilon _{IR}^2}}\left({\displaystyle \frac{m_2^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$
$`+{\displaystyle \frac{1}{\epsilon _{IR}^2}}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$
$`+2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)+2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)`$
$`+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{3m}\right]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{3m}\right]`$
$`2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(m_2^2+\mathrm{i}ฯต)f^{3m}\right]`$
$`2\text{Li}_2[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{3m}]\}+๐ช(ฯต_{IR}),`$
the adjacent (โhardโ) two$``$mass scalar box integral is
$`I_4^{2mh}(s,t;m_3^2,m_4^2)={\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{r_\mathrm{\Gamma }}{st}}\{{\displaystyle \frac{2}{\epsilon _{IR}^2}}[\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$ (70)
$`+\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}]`$
$`+{\displaystyle \frac{1}{\epsilon _{IR}^2}}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$
$`+2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{2mh}\right]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{2mh}\right]`$
$`2\text{Li}_2[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{2mh}]\}+๐ช(ฯต_{IR}),`$
the opposite (โeasyโ) two$``$mass scalar box integral is
$`I_4^{2me}(s,t;m_2^2,m_4^2)={\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{r_\mathrm{\Gamma }}{stm_2^2m_4^2}}`$ (71)
$`\{{\displaystyle \frac{2}{\epsilon _{IR}^2}}[\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}+\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_2^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$
$`\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}]+2\text{Li}_2[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{2me}]`$
$`+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{2me}\right]2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(m_2^2+\mathrm{i}ฯต)f^{2me}\right]`$
$`2\text{Li}_2[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{2me}]\}+๐ช(ฯต_{IR}).`$
An essential feature of the above expression for $`I_4^{2me}`$ is that it is well behaved in the $`m_2^20`$, $`m_4^20`$ limits. A consequence of this is that it contains both the one$``$mass scalar box integral and the massless scalar box integral. Thus, setting $`m_2^2=0`$ in (71) and making use of the fact that $`\mathrm{Li}_2(1)`$=$`\pi ^2/6`$, we find that the one$``$mass box scalar integral is
$`I_4^{1m}(s,t;m_4^2)={\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{r_\mathrm{\Gamma }}{st}}\{{\displaystyle \frac{2}{\epsilon _{IR}^2}}[\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$ (72)
$`+\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}]`$
$`+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{1m}\right]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{1m}\right]`$
$`2\text{Li}_2[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{1m}]{\displaystyle \frac{\pi ^2}{3}}\}+๐ช(ฯต_{IR}).`$
Finally, setting $`m_4^2=0`$ in (72), we find that the massless scalar box integral is given by
$`I_4^{0m}(s,t)={\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}{\displaystyle \frac{r_\mathrm{\Gamma }}{st}}\{{\displaystyle \frac{2}{\epsilon _{IR}^2}}[\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}`$ (73)
$`+\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)^{\epsilon _{IR}}]+2\text{Li}_2[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{0m}]`$
$`+2\text{Li}_2[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{0m}]2{\displaystyle \frac{\pi ^2}{3}}\}+๐ช(ฯต_{IR}).`$
The above expressions for the one$``$loop IR divergent scalar box integrals $`I_4^K(s,t;\{m_i^2\})`$ constitute the main result of this paper. It is important to emphasize that, owing to the fact that we have kept the โcausalโ $`\mathrm{i}ฯต`$ systematically throughout the calculation, these expressions are valid for arbitrary values of the relevant kinematic variables: external masses $`m_i^2(i=2,3,4)`$ and the Mandelstam variables $`s`$ and $`t`$.
As stated in the Introduction, the integrals $`I_4^K(s,t,\{m_i^2\})`$ have been evaluated in Ref. bern with the help of the partial differential equation technique. The calculation has been performed in the Euclidean region, where all kinematic variables are negative, i.e.,
$$s,t<0,m_2^2,m_3^2,m_4^2<0,$$
(74)
and the results thus obtained have been analytically continued to the positive values of the kinematic variables (physical region) by applying the following replacements:
$$ss+\mathrm{i}ฯต,tt+\mathrm{i}ฯต,m_i^2m_i^2+\mathrm{i}ฯต.$$
(75)
In order to facilitate the comparison of our results with those of Ref. bern , we have written our results in the same form in which they have been presented in Ref. bern . A glance at the expressions (69$``$73) reveals that they all have the same general form, namely,
$`I_4^K(s,t;\{m_i^2\})`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}r_\mathrm{\Gamma }g^K[{\displaystyle \frac{G^K(s,t;\epsilon _{IR};\{m_i^2\})}{\epsilon _{IR}^2}}`$ (76)
$`+H^K(s,t;\{m_i^2\})]+๐ช(ฯต_{IR})`$
$`K\{3m,2mh,2me,1m,0m\}.`$
The IR divergences (both soft and collinear) of the integrals are contained in the first term within the square brackets, while the second term is finite. The function $`G^K(s,t;\epsilon _{IR};\{m_i^2\})`$ is represented by a sum of powerlike terms, it depends on $`\epsilon _{IR}`$ and is finite in the $`\epsilon _{IR}0`$ limit. As for the function $`H^K(s,t;\{m_i^2\})`$, it is given in terms of dilogarithmic functions and constants.
Comparing our results with the corresponding ones of Ref. bern , we find that the expressions for $`G^K(s,t;ฯต_{IR};\{m_i^2\})`$ are in agreement. On the other hand, the corresponding expressions for the terms $`H^K(s,t;\{m_i^2\})`$ are of different form. Proving the equivalence of our results for the integrals $`I_4^K(s,t;\{m_i^2\})`$ with those of Ref. bern then amounts to showing that the expressions for the terms $`H^K(s,t;\{m_i^2\})`$ agree numerically. By doing this, we have arrived at the following conclusions: First, the results are in complete agreement in the Euclidean region. Second, for the integrals $`I_4^{2mh}(s,t;m_3^2,m_4^2)`$, $`I_4^{1m}(s,t;m_4^2)`$, and $`I_4^{0m}(s,t)`$, we have found agreement for arbitrary values of the kinematic variables. Third, the results for the integrals $`I_4^{3m}(s,t;,m_2^2,m_3^2,m_4^2)`$ and $`I_4^{2me}(s,t;m_2^2,m_4^2)`$ do not agree outside the Euclidean region.
The reason for this disagreement is that the analytical continuation from the Euclidean to the physical region as given by Eq. (75) is not well defined for all terms appearing in the expressions for the integrals $`I_4^K(s,t;m_i^2)`$ of Ref. bern . This has been pointed out in Ref. binoth . Thus, applying the replacements (75), no cut will be hit by the powerlike terms, logarithms, and the dilogarithms with a single ratio of the kinematical variables. In addition to this kind of terms, the expressions for the integrals $`I_4^{3m}(s,t;,m_2^2,m_3^2,m_4^2)`$ and $`I_4^{2me}(s,t;m_2^2,m_4^2)`$, given in Ref. bern , contain terms of the form
$$\text{Li}_2\left(1\frac{m_2^2m_4^2}{st}\right),$$
(77)
i.e., the dilogarithms of a product of ratios of the kinematic variables. This type of term requires special care. In order to avoid crossing a cut, in this case one has to make the following replacements:
$`\text{Li}_2\left(1{\displaystyle \frac{m_2^2m_4^2}{st}}\right)\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+\eta ({\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}},{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})\mathrm{ln}\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right),`$
where the function $`\eta `$ is defined as
$$\eta (x,y)=\mathrm{ln}(xy)\mathrm{ln}(x)\mathrm{ln}(y),$$
(79)
and arises from the possibility that $`x`$, $`y`$, and $`xy`$ are not on the same Riemann sheet. Choosing the principal value of the logarithm such that the cut lies along the negative real axis, Eq. (79) can be written as follows:
$`\eta (x,y)`$ $`=`$ $`2\pi \mathrm{i}\{\theta (\mathrm{Im}x)\theta (\mathrm{Im}y)\theta (\mathrm{Im}xy)`$ (80)
$`\theta (\mathrm{Im}x)\theta (\mathrm{Im}y)\theta (\mathrm{Im}xy)\}.`$
If the terms of the form given in (77) are analytically continued in accordance with (3.3), we find that the results of Ref. bern for the integrals $`I_4^{3m}(s,t;m_2^2,m_3^2,m_4^2)`$ and $`I_4^{2me}(s,t;m_2^2,m_4^2)`$ are numerically equivalent to the corresponding results obtained in this paper for arbitrary values of kinematic variables.
Having thus numerically established the equivalence of the two sets of results for the integrals $`I_4^K(s,t;\{m_i^2\})`$, our next task is to demonstrate this equivalence analytically, i.e., explicitly. There are, in principle, two approaches to accomplish this. The first approach consists in applying a series of the Hill identities hill (relating the dilogarithms of different arguments) to the final expression for the integral $`Q^{3m}`$ given by Eq. (3.2), with the aim to express it in terms of the dilogarithms occuring in the final expression for the integral $`I_4^{3m}(s,t;\{m_i^2\})`$ in Ref. bern . However, because of the presence of $`\mathrm{i}ฯต`$ in the arguments of the dilogarithms in Eq. (3.2), this turns out to be extremely messy. In the second approach, one tries to achieve the same by recalculating the integral in (61) with an appropriate change of the integration variable. After a lengthy calculation, details of which are presented in Appendix A, we have been able to show that
$`Q^{3m}=\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$ (81)
$`\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right){\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{s+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+\eta ({\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}},{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})\mathrm{ln}\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+๐ช(\epsilon _{IR}).`$
Although different in form, one can readily prove that this expression for $`Q^{3m}`$ is numerically equivalent to that given in Eq. (3.2) for arbitrary values of kinematic variables. All other integrals $`Q^K,K\{2mh,2me,1m,0m\}`$ can be derived by taking appropriate zero$``$mass limits. Combining the expressions thus obtained for $`Q^K`$ with expressions (51), (52), and (54) for $`P^K`$, we arrive at the final expressions for the integrals $`I_4^K(s,t;\{m_i^2\})`$ which are in agreement with those of Ref. bern if provided the latter are correctly analytically continued outside the Euclidean region.
## 4 Conclusion
Using the Feynman parameter method, we have calculated in an elegant manner a set of one$``$loop box scalar integrals with massless internal lines, but containing 0, 1, 2, or 3 nonzero external masses. To treat IR divergences (both soft and collinear), the dimensional regularization method has been employed. We have kept the causal i$`ฯต`$ systematically throughout the calculation. Consequently, the results for these integrals, which appear in the process of evaluating one$``$loop $`(N5)`$point integrals and in subdiagrams in QCD loop calculations, have been obtained for arbitrary values of the kinematic variables and represent the extension of the results of Ref. bern outside the Euclidean region.
###### Acknowledgements.
This work was supported by the Ministry of Science and Technology of the Republic of Croatia under Contract No. 00980102.
## Appendix A
In this Appendix we analytically demonstrate that for Euclidean kinematics our results for the intgerals $`I_4^K(s,t;\{m_i^2\})`$ are in agreement with the corresponding results obtained in Ref. bern . As a byproduct, we prove that the correct analytical continuation for the terms of the form given in (77), which appear in the expressions for the integrals $`I_4^{3m}(s,t;m_2^2,m_3^2,m_4^2)`$ and $`I_4^{2me}(s,t;m_2^2,m_4^2)`$ of Ref. bern , is given by Eq. (3.3).
To accomplish that, we return to the integral $`Q^{3m}`$ given in Eq. (61), in which, for convenience, we replace the integration variable $`z`$ by $`y`$. Passing to the new integration variable given by
$$z=1\frac{y(tm_2^2)t\mathrm{i}ฯต}{y(m_4^2s)m_4^2\mathrm{i}ฯต},$$
(82)
the integral $`Q^{3m}`$ becomes a line integral
$`Q^{3m}`$ $`=`$ $`{\displaystyle _{z_1}^{z_2}}{\displaystyle \frac{\mathrm{d}z}{z}}{\displaystyle \frac{s+tm_2^2m_4^2}{z(m_4^2s)+s+tm_2^2m_4^2}}\mathrm{ln}(1z)`$ (83)
$`+๐ช(\epsilon _{IR}),`$
with the path of integration followed from $`z_1`$ to $`z_2`$, where
$$z_1=1\frac{t+\mathrm{i}ฯต}{m_4^2+\mathrm{i}ฯต},z_2=1\frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}$$
(84)
Applying the partial fraction decomposition (23), this integral can be represented as
$$Q^{3m}=I+J+๐ช(\epsilon _{IR}),$$
(85)
where
$$I=_{z_1}^{z_2}\frac{\mathrm{d}z}{z}\mathrm{ln}(1z)$$
(86)
and
$$J=_{z_1}^{z_2}dz\frac{m_4^2s}{z(m_4^2s)+s+tm_2^2m_4^2}\mathrm{ln}(1z).$$
(87)
Let us now consider these two integrals in turn.
The integrand in (86) has a first$``$order pole at $`z=0`$ and the logarithmic branch cut extending from 1 to $`\mathrm{}`$. It follows from Eq. (84) that, depending on the values of the parameters $`s,t,m_2^2,m_4^2`$, it might happen that the line connecting the points $`z_1`$ and $`z_2`$ crosses the real axis. In this case, as it can be seen from Eq. (82), the crossing occurs at the point $`z=0`$ $``$ the pole of the integrand. Observe, however, that the residue of the integrand at $`z=0`$ is zero. A consequence of this is that, regardless of whether the line connecting $`z_1`$ and $`z_2`$ crosses the real axis or not, we are allowed to assume that it passes through the point $`z=0`$. Consequently, the line between $`z_1`$ and $`z_2`$ can be decomposed into two segments: one connecting $`z_1`$ and 0, and the other connecting $`0`$ and $`z_2`$. In view of this, the integral $`I`$ can be rewritten in the form
$$I=_0^{z_2}\frac{\mathrm{d}z}{z}\mathrm{ln}(1z)_0^{z_1}\frac{\mathrm{d}z}{z}\mathrm{ln}(1z).$$
(88)
By changing the integration variables $`zzz_2`$ and $`zzz_1`$ in the first and second integral, respectively, and taking formula (47) into account, we obtain the result
$$I=\text{Li}_2(1\frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต})+\text{Li}_2(1\frac{t+\mathrm{i}ฯต}{m_4^2+\mathrm{i}ฯต})$$
(89)
Applying the transformation
$$\text{Li}_2\left(1\frac{1}{y}\right)=\text{Li}_2(1y)\frac{1}{2}\mathrm{ln}^2y$$
(90)
to the second term on the right$``$hand side in (89), the final expression for the integral $`I`$ is found to be
$`I`$ $`=`$ $`\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$ (91)
$`{\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
Turning now to the integral $`J`$ in (87), we switch back to the old integration variable as a result of which the integral takes the form
$$J=J_1+J_2+J_3,$$
(92)
where
$`J_1`$ $`=`$ $`{\displaystyle _0^1}dy{\displaystyle \frac{m_4^2s}{y(m_4^2s)m_4^2\mathrm{i}ฯต}}`$ (93)
$`\times \mathrm{ln}[y(m_4^2s)m_4^2\mathrm{i}ฯต],`$
$`J_2`$ $`=`$ $`{\displaystyle _0^1}dy{\displaystyle \frac{m_4^2s}{y(m_4^2s)m_4^2\mathrm{i}ฯต}}\mathrm{ln}(t\mathrm{i}ฯต),`$ (94)
and
$$J_3=_0^1\mathrm{d}y\frac{m_4^2s}{y(m_4^2s)m_4^2\mathrm{i}ฯต}\mathrm{ln}(1y\frac{tm_2^2}{t+\mathrm{i}ฯต})$$
(95)
The integrals $`J_1`$ and $`J_2`$ are elementary, and are readily evaluated. The results are
$`J_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[\mathrm{ln}^2(s\mathrm{i}ฯต)\mathrm{ln}^2(m_4^2\mathrm{i}ฯต)\right],`$ (96)
$`J_2`$ $`=`$ $`\mathrm{ln}(t\mathrm{i}ฯต)\mathrm{ln}\left({\displaystyle \frac{s+\mathrm{i}ฯต}{m_4^2+\mathrm{i}ฯต}}\right)`$ (97)
In order to evaluate the integral $`J_3`$, we proceed by adding and subtracting the following expression:
$$_0^1\frac{\mathrm{d}y}{y}\mathrm{ln}\left(1y\frac{tm_2^2}{t+\mathrm{i}ฯต}\right)+_0^1\frac{\mathrm{d}y}{y}\mathrm{ln}(1y)$$
$$+_0^1dy\frac{1}{y{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{m_4^2s}}}\mathrm{ln}(1y).$$
After a simple algebraic reduction, the integral $`J_3`$ can be represented in the form
$$J_3=J_{3,1}+J_{3,2},$$
(98)
where
$`J_{3,1}={\displaystyle _0^1}{\displaystyle \frac{\mathrm{d}y}{y}}\mathrm{ln}\left(1y{\displaystyle \frac{tm_2^2}{t+\mathrm{i}ฯต}}\right){\displaystyle _0^1}{\displaystyle \frac{\mathrm{d}y}{y}}\mathrm{ln}(1y)`$ (99)
$`+{\displaystyle _0^1}dy{\displaystyle \frac{1}{y{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{m_4^2s}}}}\mathrm{ln}(1y)`$
$`=`$ $`\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)+{\displaystyle \frac{\pi ^2}{6}}\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right),`$
and
$$J_{3,2}=_0^1dy\left[\frac{1}{y{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{m_4^2s}}}\frac{1}{y}\right]\mathrm{ln}\left(\frac{1y}{1y{\displaystyle \frac{tm_2^2}{t+\mathrm{i}ฯต}}}\right).$$
(100)
Next, consider the integral $`J_{3,2}`$. Introducing a new integration variable
$$z=\frac{1y}{1y{\displaystyle \frac{tm_2^2}{t+\mathrm{i}ฯต}}},$$
(101)
and the notation
$$a=\frac{(m_2^2+\mathrm{i}ฯต)(m_4^2+\mathrm{i}ฯต)}{(s+\mathrm{i}ฯต)(t+\mathrm{i}ฯต)},$$
(102)
the integral can be cast into the form
$$J_{3,2}=_0^1\frac{\mathrm{d}z}{z}\mathrm{ln}(1z)+_0^1dz\frac{1}{z{\displaystyle \frac{1}{1a}}}\mathrm{ln}z.$$
(103)
Notice that, in accordance with (101), the integral $`J_{3,2}`$ is given by Eq. (103) as a line integral in the complex $`z`$plane with the integration path having only the end$``$points on the real axis at $`z=0`$ and $`z=1`$. Adding and subtracting the integral of the form
$$_0^1dz\frac{1}{z{\displaystyle \frac{1}{1a}}}\mathrm{ln}(1a)$$
in the second integral in (103) allows us to write the integral $`J_{3,2}`$ as
$`J_{3,2}`$ $`=`$ $`{\displaystyle _0^1}{\displaystyle \frac{\mathrm{d}z}{z}}\mathrm{ln}(1z)+{\displaystyle _0^1}dz{\displaystyle \frac{1}{z{\displaystyle \frac{1}{1a}}}}[\mathrm{ln}z+\mathrm{ln}(1a)]`$ (104)
$`{\displaystyle _0^1}dz{\displaystyle \frac{1}{z{\displaystyle \frac{1}{1a}}}}\mathrm{ln}(1a).`$
It should be observed that the residue of the first integral on the right$``$hand side in (101) at the pole $`z=0`$ is equal to zero. The same is true for the second integral at the pole $`z=1/(1a)`$. Therefore, the integration path in both of these integrals can be taken to follow the real axis from 0 to 1. After evaluating the first two integrals in (104) and passing to the old integration variable in the third integral, we arrive at the following expression for $`J_{3,2}`$:
$`J_{3,2}={\displaystyle \frac{\pi ^2}{6}}+\text{Li}_2(1a)+\mathrm{ln}(a)\mathrm{ln}(1a)+\mathrm{ln}(1a)`$ (105)
$`\times {\displaystyle _0^1}\mathrm{d}y{\displaystyle \frac{stm_2^2m_4^2+\mathrm{i}ฯต(s+tm_2^2m_4^2)}{[y(m_4^2s)m_4^2\mathrm{i}ฯต][y(tm_2^2)t\mathrm{i}ฯต]}}.`$
Carrying out the remaining integration, we obtain the final expression
$$J_{3,2}=\frac{\pi ^2}{6}+\text{Li}_2(1a)+\eta (\frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต},\frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต})\mathrm{ln}(1a),$$
(106)
where the function $`\eta (x,y)`$ is defined by Eq. (79). Now, substituting (99) and (106) into (98), we find the integral $`J_3`$ to be given by
$`J_3`$ $`=`$ $`\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)`$
$`+\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+\eta ({\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}},{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})\mathrm{ln}(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})`$
Finally, having obtained all the necessary ingredients, we now combine them to obtain (81), which is the desired result.
## Appendix B
By expanding the powerlike terms appearing in Eqs. (69)$``$ (73), and the prefactor $`r_\mathrm{\Gamma }`$ defined by (68), we find that the integrals under consideration can be written in the generic form
$`I_4^K`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{(4\pi )^2}}g^K\left({\displaystyle \frac{A^K}{\epsilon _{IR}^2}}+{\displaystyle \frac{B^K}{\epsilon _{IR}}}+C_1^K+C_2^K\right)+๐ช(\epsilon _{IR}),`$ (108)
$`K\{3m,2mh,2me,1m,0m\}`$
The functions $`g^K`$ appearing above are defined by (29). For convenience, the finite parts have been decomposed into two terms where the term $`C_1^K`$ originates from the expansion of the product of the $`r_\mathrm{\Gamma }`$ prefactor with the powerlike terms.
The double$``$pole parts $`A^K`$, the single$``$pole parts $`B^K`$, as well as the finite parts $`C_1^K`$ and $`C_2^K`$ of the individual integrals are listed below.
$`A^{3m}`$ $`=`$ $`0,`$
$`B^{3m}`$ $`=`$ $`\mathrm{ln}\left({\displaystyle \frac{s+\mathrm{i}ฯต}{m_2^2+\mathrm{i}ฯต}}\right)+\mathrm{ln}\left({\displaystyle \frac{t+\mathrm{i}ฯต}{m_4^2+\mathrm{i}ฯต}}\right),`$
$`C_1^{3m}`$ $`=`$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2+\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2`$
$``$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_2^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2`$
$``$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2`$
$`+`$ $`{\displaystyle \frac{1}{2}}\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)+\mathrm{ln}\left({\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)\right]^2`$
$`+`$ $`{\displaystyle \frac{1}{2}}\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)+\mathrm{ln}\left({\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)\right]^2,`$
$`C_2^{3m}`$ $`=`$ $`2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)+2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)`$
$`+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{3m}\right]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{3m}\right]`$
$`2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(m_2^2+\mathrm{i}ฯต)f^{3m}\right]2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{3m}\right].`$
$`A^{2mh}`$ $`=`$ $`1,`$
$`B^{2mh}`$ $`=`$ $`\gamma _E+\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)+\mathrm{ln}\left({\displaystyle \frac{s+\mathrm{i}ฯต}{m_4^2+\mathrm{i}ฯต}}\right)`$
$`+`$ $`\mathrm{ln}\left({\displaystyle \frac{t+\mathrm{i}ฯต}{m_3^2+\mathrm{i}ฯต}}\right),`$
$`C_1^{2mh}`$ $`=`$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2+\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2`$
$`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2`$
$`+{\displaystyle \frac{1}{2}}\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_3^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)+\mathrm{ln}\left({\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)\right]^2{\displaystyle \frac{\pi ^2}{12}},`$
$`C_2^{2mh}`$ $`=`$ $`2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)2\text{Li}_2\left(\mathrm{\hspace{0.17em}1}{\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$ (110)
$`+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{2mh}\right]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{2mh}\right]`$
$`2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{2mh}\right].`$
$`A^{2me}`$ $`=`$ $`0,`$
$`B^{2me}`$ $`=`$ $`2\mathrm{ln}\left({\displaystyle \frac{s+\mathrm{i}ฯต}{m_2^2+\mathrm{i}ฯต}}\right)+2\mathrm{ln}\left({\displaystyle \frac{t+\mathrm{i}ฯต}{m_4^2+\mathrm{i}ฯต}}\right),`$
$`C_1^{2me}`$ $`=`$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2+\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2`$
$`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_2^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2,`$
$`C_2^{2me}`$ $`=`$ $`2\text{Li}_2[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{2me}]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{2me}\right]`$
$`2\text{Li}_2[\mathrm{\hspace{0.17em}1}(m_2^2+\mathrm{i}ฯต)f^{2me}]2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{2me}\right].`$
$`A^{1m}`$ $`=`$ $`2,`$
$`B^{1m}`$ $`=`$ $`2\gamma _E+2\mathrm{ln}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)+2\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)`$
$``$ $`2\mathrm{ln}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right),`$
$`C_1^{1m}`$ $`=`$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2+\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2`$
$``$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{m_4^2\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2{\displaystyle \frac{\pi ^2}{6}},`$
$`C_2^{1m}`$ $`=`$ $`2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{1m}\right]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{1m}\right]`$ (112)
$``$ $`2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(m_4^2+\mathrm{i}ฯต)f^{1m}\right]{\displaystyle \frac{\pi ^2}{3}}.`$
$`A^{0m}`$ $`=`$ $`4,`$
$`B^{0m}`$ $`=`$ $`4\gamma _E+2\mathrm{ln}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)+2\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right),`$
$`C_1^{0m}`$ $`=`$ $`\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{s\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2+\left[\gamma _E+\mathrm{ln}\left({\displaystyle \frac{t\mathrm{i}ฯต}{4\pi \mu ^2}}\right)\right]^2{\displaystyle \frac{\pi ^2}{3}},`$
$`C_2^{0m}`$ $`=`$ $`2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(s+\mathrm{i}ฯต)f^{0m}\right]+2\text{Li}_2\left[\mathrm{\hspace{0.17em}1}(t+\mathrm{i}ฯต)f^{0m}\right]`$ (113)
$``$ $`2{\displaystyle \frac{\pi ^2}{3}}.`$
In the above expressions, $`\gamma _E`$ = 0.5722 is the Euler constant.
The pole and finite parts for the integrals $`I_4^K(s,t;m_i^2)`$ obtained in Ref. bern , but with the correct analytical continuation outside the Euclidean region for the integrals $`I_4^{3m}(s,t;m_2^2,m_3^2,m_4^2)`$ and $`I_4^{2me}(s,t;m_2^2,m_4^2)`$, can be obtained simply by replacing the terms $`C_2^K(s,t;m_i^2)`$ given above by the corresponding values listed below:
$`C_2^{3m}`$ $`=`$ $`\mathrm{ln}^2\left({\displaystyle \frac{s+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`2\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)2\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+2\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+2\eta ({\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}},{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})\mathrm{ln}\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right),`$
$`C_2^{2mh}`$ $`=`$ $`\mathrm{ln}^2\left({\displaystyle \frac{s+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`2\text{Li}_2\left(1{\displaystyle \frac{m_3^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)2\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right),`$
$`C_2^{2me}`$ $`=`$ $`\mathrm{ln}^2\left({\displaystyle \frac{s+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`2\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)2\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`2\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)2\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+2\text{Li}_2\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)`$
$`+2\eta ({\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}},{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}})\mathrm{ln}\left(1{\displaystyle \frac{m_2^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right),`$
$`C_2^{1m}`$ $`=`$ $`\mathrm{ln}^2\left({\displaystyle \frac{s+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right){\displaystyle \frac{\pi ^2}{3}}`$ (117)
$`2\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{s+\mathrm{i}ฯต}}\right)2\text{Li}_2\left(1{\displaystyle \frac{m_4^2+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right),`$
$`C_2^{0m}`$ $`=`$ $`\mathrm{ln}^2\left({\displaystyle \frac{s+\mathrm{i}ฯต}{t+\mathrm{i}ฯต}}\right)\pi ^2.`$ (118)
As it is readily seen from the above expressions, the integrals $`I_4^K(s,t;\{m_i^2\})`$ are real in the Euclidean region . Outside of this region, however, the integrals acquire an imaginary part. Being given in terms of logarithms and dilogarithms, their imaginary parts can be easily determined.
Thus, with the usual definition of the logarithms on the branch cut $`\mathrm{}<\mathrm{z}0`$, one has
$$\mathrm{ln}(y\pm iฯต)=\mathrm{ln}|y|\pm \mathrm{i}\pi \theta (y).$$
(119)
Next, as it is seen from Eq. (47), the function $`\mathrm{Li}_2(y)`$ develops an imaginary part for $`y1`$, and
$$\mathrm{Re}\mathrm{Li}_2(y\pm \mathrm{i}ฯต)=\mathrm{Li}_2\left(\frac{1}{y}\right)\frac{1}{2}\mathrm{ln}^2y+\frac{\pi ^2}{3},$$
(120)
$$\mathrm{Im}\mathrm{Li}_2(y\pm \mathrm{i}ฯต)=\pm \pi \mathrm{ln}y.$$
(121)
|
warning/0006/math0006147.html
|
ar5iv
|
text
|
# Contents
## 1 Introduction
This paper is a follow-up to our previous paper , where we presented an algebraic construction of the chiral effective action for Conformal Field Theory on higher genus Riemann surfaces. The aim of the present work is two-fold.
First, in light of the renewed interest for Classical Field Theory , we present a case study for an action functional whose construction exhibits non-trivial algebraic properties โ the action is actually the evaluation of a certain Deligne class. The functional is non-topological, which should be contrasted with cases where methods of homological algebra and algebraic topology were used to construct topological terms . Furthermore, in the recent development of String Theory, there appear dynamical fields of a new geometric content, such as, for example, the $`B`$-field. It is very important to find adequate geometric structures to describe these fields and to devise suitable action functionals . Some attempts have been made at introducing the language of *gerbes* as the proper geometric structure, at least in the lower degrees (where the language itself makes sense). In this approach, one usually settles for a ฤech description relative to some open covering of the underlying manifold. Therefore an added motivation to our work, although we mention gerbes only in passing, was to show the universal nature of the ฤech paradigm for constructing action functionals. By this we mean to develop a method which works for general ฤech resolutions and cohomology with respect to *arbitrary coverings*, and not just the standard open cover, and which allows to freely change among the coverings.
This brings us to the second goal: to describe explicitly the dependence of the chiral action functional on various default choices, which is necessary in order to make our construction in work for arbitrary coverings. In particular, this calls for the following.
1. A detailed analysis of the descent equations with respect to the nerve of the cover, where the use of Deligne complexes becomes crucial.
2. An analysis of the dependence of the chiral action on the choice of the projective structure on the Riemann surface.
Recall that the choice of the universal cover for a Riemann surface, made in , yields a default choice for the projective structure: the Fuchsian projective structure, provided by the uniformization map. Since the universal Conformal Ward Identity (CWI) determines the chiral action only up to a holomorphic projective connection, the dependence of the chiral action functional on the choice of a projective structure should be compatible with it. Indeed, we prove this for the chiral action โon shellโ, i.e., for solutions of the classical equations of motion.
In order to describe the content of this paper in more detail, we briefly recall the main results in .
Let $`\mu `$ be a Beltrami coefficient on $``$ โ a smooth bounded function $`\mu `$ with the property $`\mu _{\mathrm{}}=sup_z|\mu (z)|<1`$ โ and let $`f`$ be a solution of the Beltrami equation
$$f_{\overline{z}}=\mu f_z,$$
a self-map $`f:`$, unique up to post-composition with a Mรถbius transformation. The Euclidean version of Polyakovโs action functional for two-dimensional quantum gravity has the form
$$S[f]=2\pi i_{}\frac{f_{zz}}{f_z}\mu _z๐zd\overline{z},$$
and solves the universal Conformal Ward Identity
$$(\overline{}\mu 2\mu _z)\frac{๐นW}{๐น\mu (z)}=\frac{c}{12\pi }\mu _{zzz},$$
where $`W[\mu ]`$ is the generating functional for the vacuum chiral conformal block, and
$$W[\mu ]=\frac{c}{96\pi ^2}S[f].$$
Here $`c`$ is the central charge of the theory, and we denoted by $`๐น`$ the variational operator.
In , we extended Polyakovโs ansatz from $``$ to a compact Riemann surface $`X`$ of genus $`g>1`$, using the following construction. Consider the universal cover $`X`$, where $``$ is the upper half-plane, and let $`\mu `$ be a Beltrami coefficient on $``$, which is a pull-back of a Beltrami coefficient on $`X`$ (see 2.1 and , and also for details). Depending on the extension of $`\mu `$ into the lower half-plane, there exists a unique solution $`f`$ to the Beltrami equation on $``$. It is a map $`f:๐ป`$ with the following intertwining property:
$$f\mathrm{\Gamma }=\stackrel{~}{\mathrm{\Gamma }}f,$$
where $`\mathrm{\Gamma }`$ is a Fuchsian group uniformizing the Riemann surface $`X`$ (it is isomorphic to $`\pi _1(X)`$ as an abstract group), and $`\mathrm{\Gamma }\stackrel{~}{\mathrm{\Gamma }}`$ is an isomorphism onto a discrete subgroup of $`\mathrm{PSL}_2()`$. The domain $`๐ป=f()`$ is diffeomorphic to $``$ and can be made equal to $``$ by choosing an appropriate extension of $`\mu `$. In this way one gets a deformation map $`f:X\mathrm{\Gamma }\backslash \stackrel{~}{\mathrm{\Gamma }}\backslash ๐ป\stackrel{~}{X}`$ (which is also denoted by $`f`$) onto a new Riemann surface $`\stackrel{~}{X}`$.
The de Rham complex on $``$ is a complex of $`\mathrm{\Gamma }`$-modules for the obvious pull-back action. The basic 2-form of Polyakovโs ansatz
$$\omega [f]=\frac{f_{zz}}{f_z}\mu _zdzd\overline{z}$$
on $``$ is manifestly not invariant under the action of $`\mathrm{\Gamma }`$; this means that regarding $`\omega [f]`$ as a 0-cochain for $`\mathrm{\Gamma }`$ with values in 2-forms, its group coboundary is not zero. Nevertheless, $`\omega [f]`$ can be extended to a cocycle $`\mathrm{\Omega }[f]`$ of total degree 2 living in the double complex $`๐ข^{p,q}=C^q(\mathrm{\Gamma },\underset{ยฏ}{A}^p())`$, whose total cohomology coincides with the de Rham cohomology of $`X`$. Simple integration for the genus zero case is replaced by the evaluation over a suitable representative $`\mathrm{\Sigma }`$ of the fundamental class $`[X]`$ of $`X`$, defining
$$S[f]=\mathrm{\Omega }[f],\mathrm{\Sigma }.$$
This construction extends the definition of the chiral action to a higher genus Riemann surface $`X`$, and the functional $`S[f]`$ has the same variational properties as Polyakovโs action on the complex plane. In particular, it solves the universal CWI, the general solution being the sum of $`W[\mu ]=c/96\pi ^2S[f]`$ and an arbitrary quadratic differential, holomorphic with respect to the new complex structure on $`X`$ determined by the Beltrami differential $`\mu `$.
The main advantage of working with the universal cover $``$ is that one can use formulas from the genus zero case and simply โpush them ontoโ the double complex $`๐ข^{p,q}=C^q(\mathrm{\Gamma },\underset{ยฏ}{A}^p())`$.<sup>2</sup><sup>2</sup>2Another procedure would be to find a covariant version of everything on the base $`X`$ (cf. ), but this introduces additional โbackgroundโ structures with no direct bearing to the complex and algebro-topological structures of $`X`$. However, working with the universal cover uses several default choices, as follows.
* The groups $`\mathrm{\Gamma }`$ and $`\stackrel{~}{\mathrm{\Gamma }}`$ are discrete subgroups of $`\mathrm{PSL}_2()`$ and $`\mathrm{PSL}_2()`$ respectively, so that local sections to the covering maps $`X`$ and $`๐ป\stackrel{~}{X}`$ are projective structures subordinated to the complex structures of $`X`$ and $`\stackrel{~}{X}`$, respectively. These projective structures are inherent in the choice of $``$ as a cover, and they do not appear explicitly in the expression for the total cocycle $`\mathrm{\Omega }[f]`$.
* $`H^3(X,)=0`$ has to be invoked to close the descent equations leading from $`\omega `$ to the total cocycle $`\mathrm{\Omega }`$. This fact can be interpreted as the vanishing of an obstruction or, in other words, as an integrability property for the problem of choosing integration constants to the last descent equation. An element of arbitrariness is introduced in the explicit computation of $`\mathrm{\Omega }`$ by choosing a shift of a $``$-valued 3-cochain in this equation to turn it into ฤech coboundary.
* A specific choice of logarithm branches was made in .
The analysis of this construction shows that what we have used were not some specific features of the universal cover $`X`$, but rather its algebraic properties relative to the double complex $`๐ข^{p,q}`$: the facts that $``$ is contractible, and that $`\mathrm{\Gamma }`$ is cohomologically trivial with respect to modules of smooth forms on $``$. These are precisely the properties of a โgoodโ cover , one for which the ฤech-de Rham double complex computes cohomology groups for both theories.
As in , start with the deformation map $`f:X\stackrel{~}{X}`$, defined, say, as the solution of the Beltrami equation on $`X`$. It is natural to ask whether it is possible to carry out the same scheme as with $``$ with respect to a different cover of $`X`$, for example an ordinary open cover $`๐ฐ_X=\{U_i\}_{iI}`$ of $`X`$, with the requirement that it should allow for a change of covering morphism without changing the formalism. This is achieved by considering, for a given covering map $`UX`$ and a sheaf $`\underset{ยฏ}{F}`$, or complex of sheaves $`\underset{ยฏ}{F}^{}`$ on $`X`$, its ฤech cohomology $`\stackrel{ห}{H}^{}(UX;\underset{ยฏ}{F})`$, or hypercohomology $`\stackrel{ห}{}^{}(UX;\underset{ยฏ}{F}^{})`$, respectively. The framework of the universal cover is retrieved from the observation that group cohomology for $`\mathrm{\Gamma }`$ is ฤech cohomology for the covering $`X`$.
Our main difference from is the use of the Deligne complex instead of the simpler de Rham complex. In particular, introducing the smooth de Rham sheaves $`\underset{ยฏ}{A}_X^{}`$, we work with the Deligne complex of length 3: $`(3)_๐^{}:(3)\stackrel{\mathit{ฤฑ}}{}\underset{ยฏ}{A}_X^0\stackrel{๐}{}\underset{ยฏ}{A}_X^1\stackrel{๐}{}\underset{ยฏ}{A}_X^2`$, where $`(3)\stackrel{def}{=}(2\pi i)^3`$, and apply the same procedure as before. Namely, we form the double complex $`๐ข^{p,q}=\stackrel{ห}{C}^q(UX;(3)_๐^p)`$, localize the Polyakovโs 2-form $`\omega `$ to $`U`$ as an element of degree $`(3,0)`$ in this complex<sup>3</sup><sup>3</sup>3There is a degree shift caused by the insertion of the integers at degree zero in the Deligne complex., and perform the usual descent calculations. The latter procedure was first introduced into mathematical physics in . Specifically, we solve for elements $`\theta `$ and $`\mathrm{\Theta }`$ of degree $`(2,1)`$ and $`(1,2)`$, respectively, satisfying equations $`\stackrel{ห}{\delta }\omega =d\theta `$ and $`\stackrel{ห}{\delta }\theta =d\mathrm{\Theta }`$, with $`\stackrel{ห}{\delta }\mathrm{\Theta }(3)`$, where $`\stackrel{ห}{\delta }`$ is the ฤech coboundary operator. It is crucial that these equations are solvable due to the vanishing of the tame symbol $`(TX,TX]`$ in *holomorphic* Deligne cohomology. As a result, starting from Polyakovโs 2-form $`\omega [f]`$ we obtain a cocycle $`\mathrm{\Omega }[f]`$ of total degree 3 in the total complex $`\mathrm{Tot}๐ข^,`$. This constitutes the first result of the paper, Proposition 3.3.1. Note that it is convenient, for a regular open cover $`๐ฐ_X`$, to consider the most general form of the bulk term for the Polyakovโs action, given by adding a smooth projective connection $`h`$ to the local basic 2-form for genus 0:
$$\omega [f]=\frac{f_{zz}}{f_z}\mu _zdzd\overline{z}+2\mu hdzd\overline{z}.$$
Here $`z`$ is a local coordinate for $`U๐ฐ_X`$, and $`h`$ a representative in $`U`$ of a smooth projective connection on $`X`$ โ a smooth coboundary for the usual Schwarzian cocycle relative to the cover $`๐ฐ_X`$. The space $`๐ฌ(X)`$ of all such coboundaries is an affine space over the vector space of smooth quadratic differentials on $`X`$. On $``$, the pull-back of a projective connection is a quadratic differential. See sections 2.2-2.3 and 3.2-3.3 for details.
In section 3.4, we translated the generalized ฤech formalism for the universal cover $`X`$ into group cohomology for $`\mathrm{\Gamma }\pi _1(X)`$, so that Proposition 3.3.1 translates into Proposition 3.4.1, thus refining the corresponding results in .
For the construction of the action functional, we need to evaluate the cocycle $`\mathrm{\Omega }[f]`$ against the fundamental class $`[X]`$ of a Riemann surface $`X`$, which we represent as a cycle $`\mathrm{\Sigma }`$ in a homological double complex $`๐ฒ_{p,q}=S_p(N_qU)`$ of singular $`p`$-simplices in the $`q+1`$-fold product of $`U`$ with itself. Using the pairing $`,`$ between Deligne cocycles and cycles, which is well-defined because $`dimX=2=31`$, we can define
$$S[f]=\mathrm{\Omega }[f],{}_{}{}^{\mathrm{}}\mathrm{\Sigma },$$
where $`{}_{}{}^{\mathrm{}}\mathrm{\Sigma }`$ is the shift of the cycle $`\mathrm{\Sigma }`$ so that it has total homological degree 3. Due to the insertion of integers into the Deligne complex, the pairing $`,`$ is well defined only modulo $`(3)`$, so that the action functional $`S[f]`$ is well-defined only modulo $`(3)`$. Using the exponential map $`z\mathrm{exp}\{z/(2\pi i)^2\}`$, that identifies $`/(3)`$ with $`^{}`$, one can replace the complex $`(3)_๐^{}`$ with $`\underset{ยฏ}{A}_X^{}\stackrel{d\mathrm{log}}{}\underset{ยฏ}{A}_X^1\stackrel{๐}{}\underset{ยฏ}{A}_X^2`$ and resets all degrees by one, so that cocycle $`\mathrm{\Omega }`$ would correspond to a cocycle $`\mathrm{\Psi }`$ of total degree 2. The corresponding pairing $`,_m`$ will be now multiplicative and single-valued, with values in $`^{}`$. As a result, the single-valued functional
$$A[f]=\mathrm{\Psi }[f],\mathrm{\Sigma }_m=\mathrm{exp}\{S[f]/(2\pi i)^2\}$$
is the exponential of the action, which is quite natural since we are dealing with an effective action in QFT. Details of this construction are presented in sections 2.2 and 2.4.
In section 3.3.3 we prove the independence of the the functional $`A[f]`$ from the choices of logarithm branches, establish its relations with Bloch dilogarithms, and show that it can be considered as $`^{}`$-torsor.
The second result of the paper should be understood from the view-point of Classical Field Theory. Let $`(X)๐ฏ(X)`$ be the Earle-Eells principal fibration over the Techmรผller space $`๐ฏ(X)`$. The total space $`(X)`$ of this fibration is the unit ball in the $`L^{\mathrm{}}`$ norm in the space of all smooth Beltrami differentials on $`X`$. To every $`\mu (X)`$ there corresponds a deformation map $`f(\mu ):X\stackrel{~}{X}`$, a solution of the Beltrami equation on $`X`$, uniquely determined by the condition that when pulled back to the universal cover $``$, it gives a Fuchsian deformation, i.e. $`f()=`$. This allows to consider the functional $`A[f]`$ as a map $`A:๐ฌ(X)\times (X)^{}`$.
When studying the variational problem for the functionals $`S[f]`$ and $`A[f]`$, we consider the deformation map $`f`$ as the dynamical field, and the projective connection $`h`$ as an external field, with the problem to compute the variation with respect to $`f`$. Geometrically, these variations are tangent vectors to $`(X)`$, and are of two types, depending on whether they deform the complex structure of $`X`$ or not, i.e., whether the associated Kodaira-Spencer cocycle (see section 4.1) is holomorphically trivial or not. In the former case, the variations correspond to vertical tangent vectors to the Earle-Eells fibration $`(X)๐ฏ(X)`$, and here we consider only these variations.
One needs to show that this variational problem is well-defined even though the action itself is not expressed in terms of a simple integration over $`X`$ of a 2-form. In โphysicalโ terminology, the bulk term given by the 2-form $`\omega `$ is a multi-valued one, and we prove in Theorem 4.1.1 that the variation of the action depends solely on the variation of the bulk term and is a well-defined 2-form on $`X`$. We give two proofs of this result. The first one is based on a careful analysis of the descent equations for the variations of all components of the Deligne cocycle $`\mathrm{\Omega }[f]`$. The second proof, albeit in a sketchy form, shows that this result is, in fact, more general, and depends only on descent properties of the variational bicomplex. Takensโ results are essential in this context. We plan to return to this result with more details in a more general situation, not limited to dimension 2, elsewhere.
However, this result holds only thanks to the good gluing properties of the variations, which follow from the triviality of the Kodaira-Spencer cocycle, and this formalism can not be directly applied to the case of general variations. In this respect, we point out that there was an error in the computation of general variation in the universal cover formalism . While a brute-force calculation would achieve the goal, we prefer to defer it until the development of the proper treatment of the variational formalism for multi-valued actions, where variational bicomplex(es) glue in a more complicated way due to the non-vanishing of the deformation class.
Returning to the present paper, we also give a geometric interpretation of Theorem 4.1.1. It states that at critical points under vertical variations of the dynamical field $`f`$, the external field โ the smooth projective connection $`h`$ โ is holomorphic with respect to the complex structure on $`X`$ defined by the deformation map $`f`$. In section 4.2, we reformulate this by saying that the space of critical points coincides with the pull-back to $`(X)`$ of Hubbardโs universal projective structure $`๐ซ(X)๐ฏ(X)`$, studied in .
The paper is organized as follows. In section 2 we set up some necessary tools. In particular, we give a brief tour of Deligne complexes and explain the ฤech formalism with respect to a covering $`UX`$. We also present the minimum amount of formulas necessary to perform the evaluation over representatives of the fundamental class $`[X]`$. A more in-depth presentation would have led us through a rather long detour from the main line of the paper, therefore we provide it in the appendix, in A.2. Sections 3 and 4 comprise the main body of the paper. After some general remarks in 3.2 and 3.3, we construct the representative cocycle $`\mathrm{\Omega }[f]`$, using ฤech formalism with respect to an open cover. We analyze the changes under redefinition of the logarithm branches and of the trivializing coboundary for the tame symbol $`(T_X,T_X]`$ in 3.3.5. In 3.4, we present our construction in the form suitable for coverings $`UX`$ other than the open one $`๐ฐ_X`$, and in particular translate everything in terms of $`U=`$. Finally, in 4.1 we discuss the variation of the action. After a brief reminder of some basic notions about families of projective structures in 4.2, we present in 4.3 a geometric interpretation of the vertical variation of the action.
## 2 Preliminaries and notations
### 2.1 Quasi-conformal maps and deformations
Let $`X`$ be compact Riemann surface of genus $`g>1`$. A Riemann surface is called marked, if a system of standard generators of its fundamental group $`\pi _1(X)`$ is chosen (up to an inner automorphism). Let $`๐ฏ(X)`$ be the Teichmรผller space of marked compact Riemann surfaces of genus $`g`$, with base point the Riemann surface $`X`$. It is defined as the set of equivalence classes of orientation preserving diffeomorphisms
$$f:X\stackrel{~}{X},$$
where the triples $`[X,f_1,\stackrel{~}{X}_1]`$ and $`[X,f_2,\stackrel{~}{X}_2]`$ are said to be equivalent if the map $`f_2f_1^1`$ is homotopic to a conformal mapping of $`\stackrel{~}{X}_1`$ onto $`\stackrel{~}{X}_2`$. It is well-known (see, e.g., ), that $`๐ฏ(X)`$ is a smooth manifold of real dimension $`6g6`$, and it admits a complex structure.
For any quasi-conformal map $`f:X\stackrel{~}{X}`$, let $`\mu =\mu (f)`$ be the Beltrami differential for $`X`$ associated to $`f`$. It is a section of $`TX\overline{T}X^{}`$, where $`TX`$ is the *holomorphic* tangent bundle of $`X`$, satisfying the Beltrami equation
$$\overline{}f=\mu f,$$
where $`=/z,\overline{}=/\overline{z}`$. Conversely, if a $`C^{\mathrm{}}`$ Beltrami differential $`\mu `$ has $`L^{\mathrm{}}`$-norm less than one, $`\mu _{\mathrm{}}<1`$, then the Beltrami equation is solvable and its solution $`f`$ is a diffeomorphism.
Denote by $`A^{1,1}(X)=\mathrm{\Gamma }(X,TX\overline{T}X^{})`$ the vector space of all smooth Beltrami differentials for $`X`$, and by $`(X)`$ the open unit ball in $`A^{1,1}(X)`$ with respect to the $`L^{\mathrm{}}`$-norm. It is known that $`(X)`$ is the total space of a smooth infinite-dimensional principal fibration over $`๐ฏ(X)`$ with structure group $`๐ข(X)`$, the group of all orientation preserving diffeomorphisms of $`X`$ isotopic to the identity . Briefly, for every $`\mu (X)`$ we lift it to the universal cover $``$ and consider the solution $`f(\mu )`$ of the Beltrami equation on $``$ with the condition that $`f()=`$. Such an $`f`$ exists and is unique up to a post-composition with Mรถbius automorphism of $``$. If $`g๐ข(X)`$, then $`\mu ^g\stackrel{def}{=}\mu (fg)`$.
This provides an identification between the description of the Teichmรผller space as the space of equivalence classes of the triples $`[X,f,\stackrel{~}{X}]`$ with fixed $`X`$, and as the quotient of $`(X)`$ by $`๐ข(X)`$.
For any $`\mu (X)`$ denote by $`[\mu ]`$ the corresponding element in $`๐ฏ(X)`$ and by $`f(\mu ):XX_\mu `$ the resulting deformation of $`X`$. Though actually $`X_\mu `$ depends only on the class $`[\mu ]`$, we suppress this in the notation, and whenever the element $`\mu `$ is fixed, or clear from the context, we denote $`X_\mu `$ by $`\stackrel{~}{X}`$, as above.
Let $`A^{p,q}(X)=\mathrm{\Gamma }(X,TX_{}^{}{}_{}{}^{p}\overline{T}X_{}^{}{}_{}{}^{q})`$ be the space of $`C^{\mathrm{}}`$ tensors of weight $`(p,q)`$, with the proviso that we take the tangent bundle whenever either $`p`$ or $`q`$ is negative (like $`A^{1,1}(X)`$ for Beltrami differentials). Denote by $`\underset{ยฏ}{A}_X^{p,q}`$ the corresponding sheaves of sections. It is well-known that the operator
$$\overline{}_\mu =\overline{}\mu k\mu :\underset{ยฏ}{A}_X^{k,0}\underset{ยฏ}{A}_X^{k,1}$$
(2.1.1)
is the $`\overline{}`$-operator for the complex structure determined by $`\mu `$ โ the pull-back by $`f`$ of the complex structure on $`X_\mu `$. This gives rise to the exact sequence
$$0A^{1,0}(X)\stackrel{\overline{}_\mu }{}A^{1,1}(X)H^1(X_\mu ,\mathrm{\Theta }_\mu )0,$$
where $`\mathrm{\Theta }_\mu `$ is the tangent sheaf of $`X_\mu `$, which is isomorphic to
$$0T_\mu ((X)/๐ฏ(X))T_\mu ((X))T_{[\mu ]}(๐ฏ(X))0,$$
and provides the canonical identification $`T_{[\mu ]}(๐ฏ(X))=H^1(X_\mu ,\mathrm{\Theta }_\mu )`$ (see, e.g. ).
### 2.2 Sheaves and Deligne complexes
For any smooth manifold $`M`$, we denote by $`\underset{ยฏ}{A}_M^p`$ the sheaf of smooth complex-valued $`p`$-forms on $`M`$, and by $`A^p(M)`$ the corresponding spaces of global sections. Then $`\underset{ยฏ}{A}_M^0\underset{ยฏ}{A}_M`$, the sheaf of smooth complex-valued functions. When $`M`$ is complex, we denote by $`\underset{ยฏ}{\mathrm{\Omega }}_M^p`$ the sheaves of holomorphic $`p`$-forms. In particular, $`\underset{ยฏ}{\mathrm{\Omega }}_M^0๐ช_M`$, the structure sheaf.
Recall that the *hypercohomology* groups $`^p(M,\underset{ยฏ}{F}^{})`$ of a complex of sheaves
$$\underset{ยฏ}{F}^{}:\underset{ยฏ}{F}^0\underset{ยฏ}{F}^1\mathrm{}$$
are defined as the cohomology groups of the total complex of a suitable resolution $`\underset{ยฏ}{I}^,`$ of the complex $`\underset{ยฏ}{F}^{}`$. In practice, one usually takes a ฤech resolution relative to some (sufficiently fine) cover $`๐ฐ_M`$ of $`M`$ and considers the double complex
$$๐ข^{p,q}\stackrel{def}{=}\stackrel{ห}{C}^q(๐ฐ_M,\underset{ยฏ}{F}^p).$$
The hypercohomology $`^p(M,\underset{ยฏ}{F}^{})`$ is computed by taking $`H^p(\mathrm{Tot}๐ข^,)`$, with the convention that the total differential $`D`$ in degree $`(p,q)`$ is given by $`D=d+(1)^p\stackrel{ห}{\delta }`$, where $`d`$ is the differential in the complex $`\underset{ยฏ}{F}^{}`$ and $`\stackrel{ห}{\delta }`$ is the differential in the ฤech direction. Furthermore, two complexes $`\underset{ยฏ}{F}^{}`$ and $`\underset{ยฏ}{G}^{}`$ are said to be *quasi-isomorphic* if there is a morphism $`\underset{ยฏ}{F}^{}\underset{ยฏ}{G}^{}`$ inducing an isomorphism of their cohomology sheaves: $`H^{}(\underset{ยฏ}{F})\stackrel{}{}H^{}(\underset{ยฏ}{G})`$. The standard (spectral sequence) argument implies that their hypercohomology groups are the same. We will apply this machinery to the case when the complex $`\underset{ยฏ}{F}^{}`$ is a smooth Deligne complex.
The use of Deligne complexes is nowadays fairly common, so we just recall the notations and a few basic facts needed in the sequel. It is convenient to use the โalgebraic geometersโ twistโ and set $`(p)\stackrel{def}{=}(2\pi i)^p`$. Following we have:
###### Definition 2.2.1.
Let $`M`$ be a smooth manifold. The following complex of sheaves
$$(p)_๐^{}:(p)_M\stackrel{\mathit{ฤฑ}}{}\underset{ยฏ}{A}_M\stackrel{๐}{}\underset{ยฏ}{A}_M^1\stackrel{๐}{}\mathrm{}\stackrel{๐}{}\underset{ยฏ}{A}_M^{p1}$$
is called the *smooth Deligne complex*. The *smooth Deligne cohomology groups* of $`M`$ โ denoted by $`H_๐^q(M,(p))`$ โ are the hypercohomology groups $`^q(M,(p)_๐^{})`$.
###### Remark 2.2.2.
$`(p)`$ is placed in degree zero and the degree of each term $`\underset{ยฏ}{A}_M^r`$ in $`(p)_๐^{}`$ is $`r+1`$. The first differential is just the inclusion $`ฤฑ`$ of $`(p)`$ in $`\underset{ยฏ}{A}_X`$, while $`d`$ is the usual de Rham differential. The complex is truncated to zero after degree $`p`$. An equivalent definition of the Deligne complex is presented in the appendix, cf. A.1.
The exponential map $`e:\underset{ยฏ}{A}_M\underset{ยฏ}{A}_M^{}`$, $`f\mathrm{exp}(f/(2\pi i)^{p1})`$, induces a quasi-isomorphism
$$(p)_๐^{}(\underset{ยฏ}{A}_M^{}\stackrel{d\mathrm{log}}{}\underset{ยฏ}{A}_M^1\stackrel{๐}{}\mathrm{}\stackrel{๐}{}\underset{ยฏ}{A}_M^{p1})[1],$$
where $`[1]`$ denotes the operation of shifting a complex one step to the right. Namely, for a complex $`\underset{ยฏ}{F}^{}`$ the complex $`\underset{ยฏ}{F}^{}[1]`$ is defined as $`\underset{ยฏ}{F}[1]^k=\underset{ยฏ}{F}^{k1}`$, with $`d_{[1]}=d`$.
To prove this quasi-isomorphism, observe that the non zero cohomology sheaves of the complex $`(p)_๐^{}`$ are $`_M/(p)_M`$ and $`\underset{ยฏ}{A}_M^{p1}/d\underset{ยฏ}{A}_M^{p2}`$, located in degree $`1`$ and $`p`$, respectively. Next, consider the standard exponential exact sequence $`0(p)_M\stackrel{๐}{}\underset{ยฏ}{A}_M\stackrel{๐}{}\underset{ยฏ}{A}_M^{}1`$, implying the following commutative diagram
$$\begin{array}{ccccccccc}(p)_M& \stackrel{\iota }{}& \underset{ยฏ}{A}_M& \stackrel{d}{}& \underset{ยฏ}{A}_M^1& \stackrel{d}{}& \mathrm{}& \stackrel{d}{}& \underset{ยฏ}{A}_M^{p1}\\ & & e& & & & & & & & \\ & & \underset{ยฏ}{A}_M^{}& \stackrel{d\mathrm{log}}{}& \underset{ยฏ}{A}_M^1& \stackrel{d}{}& \mathrm{}& \stackrel{d}{}& \underset{ยฏ}{A}_M^{p1}\end{array}$$
where the first vertical arrow on the left is the exponential map, and the others are given by multiplication by $`(1)^{k1}/(2\pi i)^{p1}`$ in degree $`k`$. Now it is obvious that the two complexes have the same cohomology sheaves (by identifying $`/(p)^{}`$ through the exponential map) and therefore have the same hypercohomology groups, up to an index shift: $`H_๐^q(M,(p))^{q1}(M,\underset{ยฏ}{A}_M^{}\underset{ยฏ}{A}_M^1\mathrm{}\underset{ยฏ}{A}_M^{p1})`$.
###### Remark 2.2.3.
In general, the truncation of the Deligne complex $`(p)_๐^{}`$ after degree $`p`$ is fundamental. However, when $`dimM=p1`$, this truncation is irrelevant. In other words, when the length of the complex coincides with the dimension, $`(p)_๐^{}`$ becomes an augmented de Rham complex: $`(p)_M\underset{ยฏ}{A}_M^{}`$ . Therefore the only non trivial cohomology sheaf occurs in degree 1, and $`(p)_๐^{}`$ becomes quasi-isomorphic to $`_M/(p)_M[1]`$. As a result,
$$H_๐^q(M,(p))H^{q1}(M,/(p))H^{q1}(M,^{}),$$
where the latter isomorphism is given by the exponential map.
Working out explicitly the first cohomology groups, one gets the following isomorphisms: $`H_๐^1(M,(1))H^0(M,\underset{ยฏ}{A}_M^{})`$ โ the multiplicative group of global invertible functions โ $`H_๐^2(M,(1))H^1(M,\underset{ยฏ}{A}_M^{})`$ โ the group of isomorphism classes of smooth line bundles โ and $`H_๐^2(M,(2))^1(M,\underset{ยฏ}{A}_M^{}\underset{ยฏ}{A}_M^1)`$ โ the group of isomorphism classes of line bundles with connection. Higher Deligne cohomology groups describe more complicated higher geometric structures โ e.g., *gerbes* and $`2`$-*gerbes*.
When $`M`$ is complex, there is an entirely analogous definition for the *holomorphic Deligne complex:*
$$(p)_{๐,\mathrm{โ๐๐}}^{}:(p)_M\stackrel{\mathit{ฤฑ}}{}\underset{ยฏ}{\mathrm{\Omega }}_M\stackrel{๐}{}\underset{ยฏ}{\mathrm{\Omega }}_M^1\stackrel{๐}{}\mathrm{}\stackrel{๐}{}\underset{ยฏ}{\mathrm{\Omega }}_M^{p1},$$
with the *holomorphic Deligne cohomology groups* $`H_{๐,\mathrm{โ๐๐}}^{}(M,(p))`$ being the hypercohomology groups of the complex $`(p)_{๐,\mathrm{โ๐๐}}^{}`$.
Many of the formal properties of the smooth Deligne complex are also valid in the holomorphic category. In particular, there is the exponential quasi-isomorphism
$$(p)_{๐,\mathrm{โ๐๐}}^{}(\underset{ยฏ}{\mathrm{\Omega }}_M^{}\stackrel{d\mathrm{log}}{}\underset{ยฏ}{\mathrm{\Omega }}_M^1\stackrel{๐}{}\mathrm{}\stackrel{๐}{}\underset{ยฏ}{\mathrm{\Omega }}_M^{p1})[1],$$
since non trivial cohomology sheaves of these complexes occur only in degrees $`1`$ and $`p`$ and coincide, which implies the isomorphism in the hypercohomology, so that $`H_{๐,\mathrm{โ๐๐}}^q(M,(p))^{q1}(M,\underset{ยฏ}{\mathrm{\Omega }}_M^{}\underset{ยฏ}{\mathrm{\Omega }}_M^1\mathrm{}\underset{ยฏ}{\mathrm{\Omega }}_M^{p1})`$. When $`dim_{}M=p1`$ the truncation becomes irrelevant and $`(p)_{๐,\mathrm{โ๐๐}}^{}`$ is just $`(p)_M\underset{ยฏ}{\mathrm{\Omega }}_M^{}`$. Therefore, thanks to the exactness of the holomorphic de Rham complex, $`(p)_{๐,\mathrm{โ๐๐}}^{}`$ is also quasi-isomorphic to $`_M/(p)_M[1]`$, and we have
$$^q(M,(p)_{๐,\mathrm{โ๐๐}}^{})H^{q1}(M,/(p))H^{q1}(M,^{}).$$
In particular, when $`M`$ is a Riemann surface $`X`$ and $`p=2`$ we have, for obvious dimensional reasons
$$^3(X,(2)_{๐,\mathrm{โ๐๐}}^{})H^2(X,^{})^{}$$
and
$$^4(X,(2)_{๐,\mathrm{โ๐๐}}^{})H^3(X,^{})=0.$$
These elementary facts will play a major role in the constructions in sect. 3
There is a cup product $`:(p)_๐^{}(q)_๐^{}(p+q)_๐^{}`$ given by :
$$fg=\{\begin{array}{cc}fg\hfill & \mathrm{deg}f=0,\hfill \\ fdg\hfill & \mathrm{deg}f0\text{and}\mathrm{deg}g=q,\hfill \\ 0\hfill & \text{otherwise,}\hfill \end{array}$$
and induced product in cohomology: $`:H_๐^r(M,(p))H_๐^s(M,(q))H_๐^{r+s}(M,(p+q))`$. Note that since Deligne cohomology is defined using resolutions of complexes of sheaves, one has to take into account the appropriate sign rules. That is, for two complexes $`\underset{ยฏ}{F}^{}`$ and $`\underset{ยฏ}{G}^{}`$ one forms the double complexes
$$๐ข^{p,q}(\underset{ยฏ}{F})=\stackrel{ห}{C}^q(๐ฐ_X,\underset{ยฏ}{F}^p)\text{and}๐ข^{r,s}(\underset{ยฏ}{G})=\stackrel{ห}{C}^s(๐ฐ_X,\underset{ยฏ}{G}^r)$$
and defines the cup product
$$:๐ข^{p,q}(\underset{ยฏ}{F})๐ข^{r,s}(\underset{ยฏ}{G})\stackrel{ห}{C}^{q+s}(๐ฐ_X,\underset{ยฏ}{F}^p\underset{ยฏ}{G}^r)๐ข^{p+r,q+s}(\underset{ยฏ}{F}\underset{ยฏ}{G})$$
of two elements $`\{f_{i_0,\mathrm{},i_q}\}๐ข^{p,q}(\underset{ยฏ}{F})`$ and $`\{g_{j_0,\mathrm{},j_s}\}๐ข^{r,s}(\underset{ยฏ}{G})`$ by
$$(1)^{qr}f_{i_0,\mathrm{},i_q}g_{i_q,i_{q+1},\mathrm{},i_{q+s}}.$$
In this formula, one could replace the $``$ by any other product $`\underset{ยฏ}{F}^{}\underset{ยฏ}{G}^{}(\underset{ยฏ}{F}^{}\underset{ยฏ}{G}^{})`$, in particular by the cup product for Deligne complexes, introduced above.
Brylinski and McLaughlin spell out several cup products for the first few degrees representing interesting symbol maps. We will use one of them later, so here we recall its construction.
As already observed, $`H_๐^2(M,(1))`$ corresponds to the group of smooth line bundles on $`M`$. Working out details of the ฤech resolution relative to the ฤech cover $`๐ฐ_M=\{U_i\}_{iI}`$, one finds that a class in $`H_๐^2(M,(1))`$ is represented by the cocycle $`(f_{ij},m_{ijk})`$, where $`f_{ij}\mathrm{\Gamma }(U_iU_j,\underset{ยฏ}{A}_M^0)`$ and $`m_{ijk}\mathrm{\Gamma }(U_iU_jU_k,(1)_M)`$ are subject to the relations:
$$f_{jk}f_{ik}+f_{ij}=m_{ijk},$$
$$m_{jkl}m_{ikl}m_{ijl}+m_{ijk}=0.$$
Thus $`g_{ij}=\mathrm{exp}f_{ij}`$ is a ฤech 1-cocycle with values in invertible functions, as expected.
Consider now two line bundles $`L`$ and $`L^{}`$ over $`M`$, represented by cocycles $`(f_{ij},m_{ijk})`$ and $`(f_{ij}^{},m_{ijk}^{})`$, respectively. Their cup product, to be denoted by the โtameโ symbol $`(L,L^{}]`$ (see, e.g., ), is an element of $`H_๐^4(M,(2))`$, represented by the cocycle
$$(f_{ij}df_{jk}^{},m_{ijk}f_{kl}^{},m_{ijk}m_{klp}^{}).$$
A similar interpretation holds for the holomorphic Deligne cohomology. In particular, $`H_{๐,\mathrm{โ๐๐}}^2(M,(1))`$ corresponds to the group of holomorphic line bundles on $`M`$, and the cup product of two such line bundles is $`(L,L^{}]H_{๐,\mathrm{โ๐๐}}^4(M,(2))`$. When $`dim_{}M=1`$, according to the previous remark, the cup product of two holomorphic line bundles is a trivial cocycle: $`(L,L^{}]=0`$.
In this paper our main emphasis will be on smooth Deligne cohomology in degree three. With respect to the ฤech cover $`๐ฐ_M`$, a class in $`H_๐^3(M,(3))`$ is represented by the total cocycle $`(\omega _i,a_{ij},f_{ijk},m_{ijkl})`$, where $`\omega _i\mathrm{\Gamma }(U_i,\underset{ยฏ}{A}_M^2)`$, $`a_{ij}\mathrm{\Gamma }(U_iU_j,\underset{ยฏ}{A}_M^1)`$, $`f_{ijk}\mathrm{\Gamma }(U_iU_jU_k,\underset{ยฏ}{A}_M^0)`$, and $`m_{ijkl}\mathrm{\Gamma }(U_iU_jU_kU_l,(3)_M)`$ are subject to the relations:
$`\omega _j\omega _i=da_{ij},`$ $`a_{jk}a_{ik}+a_{ij}=df_{ijk},`$ (2.2.1)
$`\stackrel{ห}{\delta }f_{ijkl}=m_{ijkl},`$ $`\stackrel{ห}{\delta }m_{ijklp}=0.`$
According to , $`H_๐^3(M,(3))`$ is the group of isomorphism classes of *gerbes* on $`M`$, equipped with *connective structure* described by $`\{a_{ij}\}`$, and with *curving* described by $`\{\omega _i\}`$.
### 2.3 ฤech formalism for generalized coverings
In this section, we provide the necessary machinery to translate statements and computations carried out in a conventional ฤech covering by open sets to other kinds of coverings, such as the universal cover, that will allow to merge results from our previous approach into the present one. This formalism is not yet part of a mathematical physics curriculum, so here we present the prerequisites necessary for computing ฤech cohomology, referring to the standard sources where the theoretical background is explained.
Let $`M`$ be a smooth manifold or topological space. The general idea is to pass from inclusions $`UM`$ to general local homeomorphisms $`UM`$ which are not necessarily injective. Technically, one fixes a category $`๐_M`$ whose objects are spaces รฉtale over $`M`$, morphisms are the covering maps, and which is closed with respect to the fiber product of the maps over $`M`$, with $`M`$ being the terminal object in $`๐_M`$. The coverings are surjective families of local homeomorphism in $`๐_M`$, namely families $`\{f_i:U_iU\}`$ of $`M`$-maps such that $`U=_if_i(U_i)`$. In practice, we shall restrict our attention to covering maps of $`M`$ itself. The key observation is that if $`U_iM`$ and $`U_jM`$ are inclusions, then $`U_iU_jU_i\times _MU_j`$ โ the fiber product of maps $`U_iM`$ and $`U_jM`$ โ so that the notion of fiber product for covering maps replaces the notion of intersection of open sets.
For a covering $`UM`$ in $`๐_M`$ we obtain an augmented simplicial object
by considering the nerve $`N_{}(UM)`$. Specifically, for any integer $`q0`$ we define
$$N_q(UM)=\underset{(q+1)\text{times}}{\underset{}{U\times _M\mathrm{}\times _MU}}$$
where for $`i=0,\mathrm{},q`$ the arrows are the maps $`d_i:N_q(UM)N_{q1}(UM)`$, forgetting the $`i\text{-th}`$ factor in the product.
For an abelian sheaf $`\underset{ยฏ}{F}`$ on $`M`$ (more precisely, on $`๐_M`$) the ฤech complex relative to a covering $`UM`$ in $`๐_M`$ is defined by setting for any $`q0`$
$$\stackrel{ห}{C}^q(U;\underset{ยฏ}{F})=\mathrm{\Gamma }(N_q(UM),\underset{ยฏ}{F})\text{with}\stackrel{ห}{\delta }=\underset{i=0}{\overset{q}{}}(1)^id_i^{}.$$
The ordinary ฤech formalism is recovered by considering an open cover $`๐ฐ_M=\{U_i\}_{iI}`$ of $`M`$ and the covering $`_{iI}U_iM`$, so that in degree $`q`$ we just get the disjoint union of all $`q`$-fold intersections
$$N_q(๐ฐ_M)=\underset{i_0,\mathrm{},i_q}{}U_{i_0}\mathrm{}U_{i_q},$$
and the resulting ฤech complex is the standard one.
At the other extreme, let $`UM`$ be a regular covering map and $`G=\mathrm{Deck}(U/M)`$ the corresponding group of deck transformations acting properly on $`U`$ on the right. One immediately verifies that
$$\underset{(q+1)\text{times}}{\underset{}{U\times _M\mathrm{}\times _MU}}U\times \underset{q\text{times}}{\underset{}{G\times \mathrm{}\times G}},$$
and under this isomorphism the maps $`d_i:N_q(UM)N_{q1}(UM)`$ become
$$d_i(x,g_1,\mathrm{},g_q)=\{\begin{array}{cc}(xg_1,g_2,\mathrm{},g_q)\hfill & i=0\hfill \\ (x,g_1,\mathrm{},g_ig_{i+1},\mathrm{},g_q)\hfill & i=1,\mathrm{},q1\hfill \\ (x,g_1,\mathrm{},g_{q1})\hfill & i=q.\hfill \end{array}$$
Hence, the ฤech complex with respect to $`UM`$ becomes the usual Eilenberg-MacLane cochain complex on $`G`$ with values in the $`G`$-module $`\underset{ยฏ}{F}(U)`$:
$$\stackrel{ห}{C}^q(U;\underset{ยฏ}{F})C^q(G;\underset{ยฏ}{F}(U)).$$
Thus the ฤech cohomology of this complex is just the group cohomology of $`G`$ with values in the $`G`$-module $`\underset{ยฏ}{F}(U)`$, where the module structure is given by the pull-back action. A particular case of special interest for us is when $`U`$ is the universal cover of $`M`$, so that $`G=\pi _1(M)`$.
The formalism clearly extends to the case where we consider a complex $`\underset{ยฏ}{A}^{}`$ of sheaves on $`M`$ โ typically, the de Rham complex. The hypercohomology with respect to a covering $`UM`$ will be the cohomology of the total complex of $`\stackrel{ห}{C}^q(U;\underset{ยฏ}{A}^p)`$.
In some favorable cases, one or both spectral sequences associated to the double complex above will degenerate at the first level. Degeneration at the first level of the first spectral sequence, that is, the one associated to the filtration on $`p`$, is equivalent to
$$\stackrel{ห}{H}^q(UM;\underset{ยฏ}{A}^p)=0\text{for all }q>0\text{.}$$
Since each $`\underset{ยฏ}{A}^p`$ is assumed to be a sheaf, that is, $`A^p(M)`$ is the kernel
$$\text{},$$
the cohomology of the total complex $`\stackrel{ห}{C}^q(U;\underset{ยฏ}{A}^p)`$ equals $`H_{\mathrm{๐๐
}}^p(M)`$.
On the other hand, the degeneration of the other spectral sequence (at the same level) means the complex $`\underset{ยฏ}{A}^{}`$ is a resolution of some sheaf $`\underset{ยฏ}{F}`$, so that the total cohomology equals $`\stackrel{ห}{H}^p(UM;\underset{ยฏ}{F})`$. Therefore, when both of these cases are realized, we have a ฤech-de Rham type situation , that is
$$^p(UM;\underset{ยฏ}{A}^{})\stackrel{ห}{H}^p(UM;\underset{ยฏ}{F})H_{\mathrm{๐๐
}}^p(M).$$
The obvious example of this situation is the ฤech-de Rham double complex relative to the ordinary cover $`_{iI}U_i`$, where the above isomorphism gives the usual de Rham theorem: $`\stackrel{ห}{H}^p(M,)H_{\mathrm{๐๐
}}^p(M)`$. Another example of utmost importance is the universal cover $`X`$ of a Riemann surface $`X`$ of genus $`g>1`$. Since there exist $`\pi _1(X)`$-equivariant partitions of unity , the sheaves $`\underset{ยฏ}{A}_{}^p`$ are acyclic: $`H^q(\pi _1(X),\underset{ยฏ}{A}_{}^p)=0`$ for $`q>0`$ and all $`p`$. Moreover, since $``$ is contractible, the de Rham complex $`\underset{ยฏ}{A}^{}()`$ is obviously acyclic in dimension greater than zero, and as a result we have the isomorphism<sup>4</sup><sup>4</sup>4See also for a simple-minded proof without spectral sequences.
$$H^p(\pi _1(X),)H_{\mathrm{๐๐
}}^p(M).$$
### 2.4 Evaluation over the fundamental class
For the construction of the action functional we need to evaluate Deligne cohomology classes against the fundamental class $`[X]`$ of $`X`$, which we need to represent as a cycle in a suitable homological double complex โ in a way analogous to the use of ฤech resolutions to compute the hypercohomology.
The aim of this section is to introduce the minimum set of tools necessary to describe the homological (double) complex and to perform the evaluation, relegating all technical details to the appendix. There, we construct an explicit representative $`\mathrm{\Sigma }`$ of $`[X]`$ with respect to a covering $`UX`$ by mirroring the cohomology computations done in 3.3. The computations are explicit enough that the reader who is only interested in the formulas for $`\mathrm{\Sigma }`$ can read A.2.3 directly. Also, the reader interested only in the construction of the local action cocycle can safely proceed to sect. 3.
As usual, whenever we mention facts that are not specific to $`X`$ being a Riemann or topological surface, we use the notation $`M`$ to denote a general smooth manifold or topological space with covering $`UM`$.
#### 2.4.1
Consider the double complex
$$๐ฒ_{p,q}=S_p(N_q(UM)),$$
where $`N_{}(UM)`$ is the nerve of the covering $`UM`$ and $`S_{}`$ is the singular simplices functor, i.e., $`S_p(M)`$ is the set of continuous maps $`\mathrm{\Delta }^pM`$, where $`\mathrm{\Delta }^p`$ is the standard simplex. For every $`p0`$, the covering map $`UM`$ induces a corresponding map $`ฯต:๐ฒ_{p,0}=S_p(U)S_p(M)`$ between simplices โ the augmentation map. The double complex $`๐ฒ_,`$ has two boundary operators: the usual boundary operator on singular chains, $`^{}:๐ฒ_{p,q}๐ฒ_{p1,q}`$, and the boundary operator $`^{\prime \prime }:๐ฒ_{p,q}๐ฒ_{p,q1}`$ induced by the face maps of the nerve: $`^{\prime \prime }=(1)^id_{i}^{}{}_{}{}^{}`$, where $`d_i:N_q(U)N_{q1}(U)`$ and $`d_{i}^{}{}_{}{}^{}`$ is the induced map on singular chains. As usual, we have the simple complex $`\mathrm{Tot}๐ฒ`$ with total differential $`=^{}+(1)^p^{\prime \prime }`$ on $`๐ฒ_{p,q}`$.
If $`U`$ is the ordinary ฤech covering $`๐ฐ_M=_{iI}U_i`$, then
$$S_p(N_q(๐ฐ_M))=\underset{i_0,\mathrm{},i_q}{}S_p(U_{i_0}\mathrm{}U_{i_q}).$$
If, on the other hand, $`U`$ is a regular covering space with $`G`$ as group of deck transformations, then $`S_p(U)`$ is a $`G`$-module with $`G`$-action given by translation of simplices. It follows that $`S_p(N_q(U))`$, for $`q>0`$, consists of simplices into $`U`$ parameterized by $`q`$-tuples of elements in $`G`$. Taking into account the expression for the face maps $`d_i`$, computed in 2.3, we get
$$S_p(N_q(U))=S_p(U)_GB_q(G),$$
where $`B_{}(G)`$ is the bar resolution and $`G`$ is the integral group ring of $`G`$. Hence, for any $`p`$, the $`^{\prime \prime }`$-homology is just the group homology
$$H_q(S_p(N_{}(U)))=H_q(G;S_p(U)).$$
We are interested in the situation when $`๐ฒ_,`$ has no homology with respect to the second index, except in degree zero, namely we want
$$H_q(S_p(N_{}(UM)))\{\begin{array}{cc}S_p(M)\hfill & q=0\hfill \\ 0\hfill & q>0,\hfill \end{array}$$
for the $`^{\prime \prime }`$ homology. In this case we say that $`๐ฒ_{p,}`$ *resolves* $`S_p(M)`$ and one has the isomorphism
$$H_{}(M,)H_{}(S_{}(M))H_{}(\mathrm{Tot}๐ฒ_,).$$
This isomorphism is induced by the augmentation map $`ฯต:\mathrm{Tot}๐ฒS_{}(M)`$, which assigns to any chain $`\mathrm{\Sigma }`$ of total degree $`n`$ in $`\mathrm{Tot}๐ฒ`$ the chain $`ฯต(\mathrm{\Sigma }_{n,0})`$, where $`\mathrm{\Sigma }_{n,0}`$ is the component in $`๐ฒ_{n,0}`$. It is easy to see that this map is a chain map, it sends cycles into cycles and induces the above isomorphism. Details can be found, e.g., in . <sup>5</sup><sup>5</sup>5A detailed calculation along these lines can also be found in the appendix of . Observe that this situation is realized for both the examples of an open ฤech cover and of a regular covering $`UM`$ (cf. the appendix). For completeness, in the appendix we briefly analyze the implications of the requirement that the double complex $`๐ฒ_,`$ is acyclic with respect to the first index, and their relations with good covers.
#### 2.4.2
For a topological manifold $`M`$ of dimension $`n`$, we need to represent $`[M]`$ with a total cycle $`\mathrm{\Sigma }`$ of degree $`n`$ in $`\mathrm{Tot}๐ฒ_,`$. It has the form
$$\mathrm{\Sigma }=\mathrm{\Sigma }_0+\underset{k=1}{\overset{n}{}}(1)^{_{l=0}^{k1}(nl)}\mathrm{\Sigma }_k,$$
where $`\mathrm{\Sigma }_k๐ฒ_{nk,k}`$ and
$$^{}\mathrm{\Sigma }_0=^{\prime \prime }\mathrm{\Sigma }_1,\mathrm{},^{}\mathrm{\Sigma }_{k1}=^{\prime \prime }\mathrm{\Sigma }_k,\mathrm{},^{}\mathrm{\Sigma }_n=0.$$
The choice of signs ensures $`\mathrm{\Sigma }=0`$, where $``$ is the total differential in $`\mathrm{Tot}๐ฒ_,`$. By definition, $`\mathrm{\Sigma }`$ is a โliftโ of $`M`$ considered as a chain in $`S_n(M)`$, i.e. $`ฯต(\mathrm{\Sigma }_0)=M`$, where $`M=_i\sigma _i`$ for a suitable collection of singular simplices $`\sigma _iS_n(M)`$. The existence of the elements $`\mathrm{\Sigma }_1,\mathrm{},\mathrm{\Sigma }_n`$ follows from the $`^{\prime \prime }`$-exactness assumption and the fact that $`\mathrm{\Sigma }_0`$ lifts $`M`$. Indeed, we have $`0=M=ฯต(\mathrm{\Sigma }_0)=ฯต(^{}\mathrm{\Sigma }_0)`$, so that there exists $`\mathrm{\Sigma }_1๐ฒ_{n1,1}`$ such that $`^{}\mathrm{\Sigma }_0=^{\prime \prime }\mathrm{\Sigma }_1`$, and so on.
Specializing to the case when $`MX`$ is a Riemann surface, the representative of the fundamental class $`[X]`$ is the cycle $`\mathrm{\Sigma }=\mathrm{\Sigma }_0+\mathrm{\Sigma }_1\mathrm{\Sigma }_2`$, with components $`\mathrm{\Sigma }_k๐ฒ_{2k,k}`$ satisfying $`^{}\mathrm{\Sigma }_0=^{\prime \prime }\mathrm{\Sigma }_1`$, $`^{}\mathrm{\Sigma }_1=^{\prime \prime }\mathrm{\Sigma }_2`$, and $`^{\prime \prime }\mathrm{\Sigma }_0=^{}\mathrm{\Sigma }_2=0`$. This cycle is explicitly constructed in the appendix for the case of an ordinary ฤech cover $`U=๐ฐ_X`$ and in for the case of the universal cover $`X`$.
Here we present the basic formulas for the ฤech case, which also gives the flavor of the general procedure which carries over to the other coverings unchanged.
Following , introduce the symbol $`\mathrm{\Delta }_{i_0,\mathrm{},i_q}`$ to denote the $`(q+1)`$-fold intersection thought as a generator in $`๐ฒ_{p,q}`$, so that a generic element can be written in the form:
$$\sigma =\underset{i_0\mathrm{}i_q}{}\sigma _{i_0\mathrm{}i_q}\mathrm{\Delta }_{i_0\mathrm{}i_q},$$
where $`\sigma _{i_0\mathrm{}i_q}`$ are are $`p`$-simplices for $`U_{i_0}\mathrm{}U_{i_q}`$, i.e. continuous maps $`\mathrm{\Delta }^pU_{i_0}\mathrm{}U_{i_q}`$. It is immediate to verify that
$$^{\prime \prime }\mathrm{\Delta }_{i_0\mathrm{}i_q}=\underset{j=0}{\overset{q}{}}(1)^j\mathrm{\Delta }_{i_0,\mathrm{},\widehat{i_j},\mathrm{},i_q},$$
where the $`\widehat{}`$ sign denotes omission. Then
$$^{\prime \prime }\sigma =\underset{i_0,\mathrm{},i_{q1}}{}\left(\underset{k=0}{\overset{q}{}}\underset{j}{}(1)^k\sigma _{i_0,\mathrm{},\underset{\stackrel{}{k\text{-th}}}{j},\mathrm{},i_q}\right)\mathrm{\Delta }_{i_0,\mathrm{},i_{q1}},$$
where the summation goes over ordered sets of indices (it is assumed that $`I`$ is an ordered set). Thus with the convention that $`_{i_0,\mathrm{},i_q}`$ is the sum over sets of indices $`\{i_0,\mathrm{},i_q\}`$ with $`i_0\mathrm{}i_q`$, we can rewrite the last equation as
$$^{\prime \prime }\sigma =\underset{k=0}{\overset{q}{}}(1)^k\underset{i_0,\mathrm{},\underset{\stackrel{}{k\text{-th}}}{j},\mathrm{},i_q}{}\sigma _{i_0,\mathrm{},j,\mathrm{},i_q}\mathrm{\Delta }_{i_0,\mathrm{},i_{q1}}.$$
Now, consider the problem of constructing the total cycle $`\mathrm{\Sigma }=\mathrm{\Sigma }_0+\mathrm{\Sigma }_1\mathrm{\Sigma }_2`$ representing $`[X]`$. Representing the components $`\mathrm{\Sigma }_i`$ as:
$$\mathrm{\Sigma }_0=\underset{i}{}\sigma _i\mathrm{\Delta }_i,\mathrm{\Sigma }_1=\underset{ij}{}\sigma _{ij}\mathrm{\Delta }_{ij},\mathrm{\Sigma }_2=\underset{ijk}{}\sigma _{ijk}\mathrm{\Delta }_{ijk},$$
we first construct $`\mathrm{\Sigma }_0`$ as follows. Starting from the nerve of the cover $`๐ฐ_X`$ consider a triangulation of $`X`$ by $`๐ฐ`$-small simplices, i.e. each simplex comprising the triangulation has support in some open set $`U_i`$ belonging to the cover (cf. the appendix for the detailed procedure). Then $`X=_i\sigma _i`$, where each chain $`\sigma _i`$ is a sum of simplices whose support is contained in $`U_i`$ for each $`i`$, and one immediately writes $`\mathrm{\Sigma }_0=_i\sigma _i\mathrm{\Delta }_i`$. The other components are determined by the $`^{\prime \prime }`$-exactness condition of the complex. Namely, from the above expression and $`^{}\mathrm{\Sigma }_0=^{\prime \prime }\mathrm{\Sigma }_1`$, $`^{}\mathrm{\Sigma }_1=^{\prime \prime }\mathrm{\Sigma }_2`$ one gets the equations
$$\underset{i}{}^{}\sigma _i\mathrm{\Delta }_i=\left(\underset{i,j}{}\sigma _{ji}\underset{i,j}{}\sigma _{ij}\right)\mathrm{\Delta }_i,$$
$$\underset{i,j}{}^{}\sigma _{ij}\mathrm{\Delta }_{ij}=\left(\underset{k,i,j}{}\sigma _{kij}\underset{i,k,j}{}\sigma _{ikj}+\underset{i,j,k}{}\sigma _{ijk}\right)\mathrm{\Delta }_{ijk}$$
for components $`\sigma _{ij}`$ and $`\sigma _{ijk}`$. Explicit expression for these components in terms of the barycentric decomposition is given in the appendix.
#### 2.4.3
In order to discuss the evaluation pairing, we need to address the issue of the index shift in the Deligne complex. One way is to explicitly use the exponential map described in 2.2 to revert the indexing to the familiar form without a shift, at the cost of introducing an explicit multiplicative structure via the exponential. Another way is to introduce an *ad hoc* index shift in homology to mirror the one in the Deligne complex, i.e. to consider singular $`q`$-simplices to be of homological degree $`q+1`$. The resulting pairing will be additive, but only defined $`mod(p)`$. The two approaches are in the end the same.
We start with second approach. Let $`(๐ช_{},)`$ be a *homological* complex. The canonical way to shift it is to introduce $`๐ช[1]_{}\stackrel{def}{=}๐ช_1`$, with $`_{[1]}=`$, cf. . We require instead that the new differential be simply $``$, while retaining the index shift. Thus we replace $`๐ฒ_{r,s}=S_r(N_s(UM))`$ by the new double complex
$${}_{}{}^{\mathrm{}}๐ฒ_{r,s}^{}=S_{r1}(N_s(UM)),$$
with differential $`=^{}+(1)^r^{\prime \prime }`$, where $`^{}`$ is the usual singular boundary, as before. If $`\mathrm{\Sigma }=(\mathrm{\Sigma }_0,\mathrm{},\mathrm{\Sigma }_q)`$, with $`\mathrm{\Sigma }_k๐ฒ_{qk,k}`$, is a $`q`$-chain in $`\mathrm{Tot}๐ฒ_,`$, then it maps to the $`(q+1)`$-chain $`{}_{}{}^{\mathrm{}}\mathrm{\Sigma }=((1)^q\mathrm{\Sigma }_0,\mathrm{},(1)^{qk}\mathrm{\Sigma }_k,\mathrm{},\mathrm{\Sigma }_q,0)`$ in $`\mathrm{Tot}{}_{}{}^{\mathrm{}}๐ฒ_{,}^{}`$, and $`{}_{}{}^{\mathrm{}}\mathrm{\Sigma }`$ is a cycle if and only if $`\mathrm{\Sigma }`$ is a cycle.
Let $`๐ข^,`$ be a ฤech resolution of the Deligne complex $`(p)_๐^{}`$ with respect to the covering $`UM`$. The pairing between $`๐ข^{r,s}`$ and $`{}_{}{}^{\mathrm{}}๐ฒ_{r,s}^{}`$ is defined as follows (cf. ). To the pair $`(\varphi ,\sigma )`$, where $`\varphi `$ is a collection $`\{\varphi _{i_0,\mathrm{},i_s}\}`$ of $`(r1)`$-forms on $`N_s(UM)`$ for $`r>0`$, or integers $`(p)`$ for $`r=0`$ and $`\sigma =\sigma _{i_0,\mathrm{},i_s}\mathrm{\Delta }_{i_0,\mathrm{},i_s}{}_{}{}^{\mathrm{}}๐ฒ_{r,s}^{}`$ we assign
$$\varphi ,\sigma =\{\begin{array}{cc}_{i_0,\mathrm{},i_s}_{\sigma _{i_0,\mathrm{},i_s}}\varphi _{i_0,\mathrm{},i_s}\hfill & r>0\hfill \\ 0\hfill & r=0.\hfill \end{array}$$
(2.4.1)
To extend this pairing to $`\mathrm{Tot}๐ข^,`$ and $`\mathrm{Tot}{}_{}{}^{\mathrm{}}๐ฒ_{,}^{}`$, let $`{}_{}{}^{\mathrm{}}\sigma =(\sigma _0,\sigma _1,\mathrm{},\sigma _{n1},0)`$, with $`\sigma _k{}_{}{}^{\mathrm{}}๐ฒ_{nk,k}^{}`$, and $`\mathrm{\Phi }=(\varphi _0,\varphi _1,\mathrm{},\varphi _n)`$, with $`\varphi _k๐ข^{nk,k}`$. Then we define
$$\mathrm{\Phi },{}_{}{}^{\mathrm{}}\sigma =\underset{k=0}{\overset{n1}{}}\varphi _k,\sigma _k,$$
(2.4.2)
where, $`\varphi _k=0`$ for all $`k<np`$, if, of course, $`n>p`$. Note that so far the pairing was defined to have values in $``$. However, the fundamental fact is that *away from the truncation degree*, i.e. when the total degree $`n`$ is strictly less than $`p`$, and therefore the form degree is strictly less than $`p1`$, the total differentials $`D`$ and $``$ are transpose to each other modulo $`(p)`$:
$$D\mathrm{\Phi },{}_{}{}^{\mathrm{}}\mathrm{\Sigma }=\mathrm{\Phi },{}_{}{}^{\mathrm{}}\mathrm{\Sigma }mod(p).$$
(2.4.3)
for all $`\mathrm{\Phi }\mathrm{Tot}๐ข^,`$, $`{}_{}{}^{\mathrm{}}\mathrm{\Sigma }\mathrm{Tot}{}_{}{}^{\mathrm{}}๐ฒ_{,}^{}`$. This readily follows from the very definition of the Deligne complex. Equation (2.4.3) means that the pairing $`,`$, *considered modulo* $`(p)`$, defines a pairing between $`H^{}(\mathrm{Tot}๐ข^,)`$ and $`H_{}(\mathrm{Tot}๐ฒ_,)`$ away from the truncation degree $`p1`$.
Formula (2.4.3) would not hold for degrees bigger or equal than $`p1`$, *unless* $`dimM=p1`$ โ the case where the truncation becomes unimportant. This is the situation we will be interested in in sect. 3. Therefore in this case the pairing (2.4.2) descends to the corresponding homology and cohomology groups and is non degenerate. It defines a pairing between $`H^{}(\mathrm{Tot}๐ข^,)`$ and $`H_{}(\mathrm{Tot}๐ฒ_,)`$ which we continue to denote by $`,`$.
Let us show how these formulas work in the case of a Riemann surface $`X`$ and a Deligne cocycle $`\mathrm{\Omega }=(\omega _i,a_{ij},f_{ijk},m_{ijkl})`$ of total degree $`3`$. (Recall that the individual elements are subjects to the relations (2.2.1).) Let $`\mathrm{\Sigma }=(\mathrm{\Sigma }_0,\mathrm{\Sigma }_1,\mathrm{\Sigma }_2)`$ be a representative in $`\mathrm{Tot}๐ฒ_,`$ of the fundamental class $`[X]`$. Then the corresponding element in the shifted complex will be
$${}_{}{}^{\mathrm{}}\mathrm{\Sigma }=(\mathrm{\Sigma }_0,\mathrm{\Sigma }_1,\mathrm{\Sigma }_2,0).$$
Omitting the indices, the evaluation of the class of $`\mathrm{\Omega }`$ over $`[X]`$ will be computed by the expression
$$\mathrm{\Omega },{}_{}{}^{\mathrm{}}\mathrm{\Sigma }=\omega ,\mathrm{\Sigma }_0a,\mathrm{\Sigma }_1f,\mathrm{\Sigma }_2,$$
(2.4.4)
where each term in (2.4.4) should be expanded according to (2.4.1). This evaluation takes its values in $`/(3)`$ and does not depend on the representative cocycle of the Deligne cohomology class $`[\mathrm{\Omega }]H_๐^3(M,(3))`$.
Another way to define the pairing is to use explicitly the quasi-isomorphism
$$(p)_๐^{}(\underset{ยฏ}{A}_M^{}\stackrel{d\mathrm{log}}{}\underset{ยฏ}{A}_M^1\stackrel{๐}{}\mathrm{}\stackrel{๐}{}\underset{ยฏ}{A}_M^{p1})[1],$$
induced by the exponential map (see 2.2). In this way a cocycle representing a class of degree $`k`$ in $`H_๐^k(M,(p))`$ becomes a cocycle of degree $`k1`$ in the double complex
$$\stackrel{ห}{C}^{}(UM,\underset{ยฏ}{A}_M^{}\underset{ยฏ}{A}_M^1\mathrm{}\underset{ยฏ}{A}_M^{p1}).$$
In particular, if $`\mathrm{\Omega }=(\omega _i,a_{ij},f_{ijk},m_{ijkl})`$, subject to the relations (2.2.1), is a cocycle of total degree 3 in $`\stackrel{ห}{C}^{}(UM,(3)_๐^{})`$ representing a Deligne class of total degree 3, the element
$$\mathrm{\Psi }=(\frac{1}{(2\pi i)^2}\omega _i,\frac{1}{(2\pi i)^2}a_{ij},\mathrm{exp}\left(\frac{f_{ijk}}{(2\pi i)^2}\right))$$
is the corresponding cocycle of total degree 2.
As in the previous discussion, we will consider only the the case when $`dimM=p1`$, where $`p`$ is the length of the Deligne complex, so that the truncation becomes irrelevant. Denote by $`\stackrel{~}{๐ข}^,`$ the double complex $`\stackrel{ห}{C}^{}(UM,\underset{ยฏ}{A}_M^{}\underset{ยฏ}{A}_M^1\mathrm{}\underset{ยฏ}{A}_M^n)`$, with $`n=p1`$.
Then there exists a natural pairing between $`\stackrel{~}{๐ข}^{r,s}`$ and $`๐ฒ_{r,s}`$ which assigns to the pair $`(\psi ,c)`$ the evaluation of the $`r`$-form $`\psi `$ over a chain $`c=\sigma _{i_0,\mathrm{},i_s}\mathrm{\Delta }_{i_0,\mathrm{},i_s}S_r(N_s(UM))`$:
$$\psi ,c=_{\sigma _{i_0,\mathrm{},i_s}}\psi _{i_0,\mathrm{},i_s},$$
with the understanding that for $`r=0`$ this is just the pointwise evaluation of an invertible function, defined through the exponential map. To define a multiplicative pairing between $`\mathrm{Tot}\stackrel{~}{๐ข}^,`$ and $`\mathrm{Tot}๐ฒ_,`$, let $`C=(c_0,c_1,\mathrm{},c_n)`$, with $`c_i๐ฒ_{ni,i}`$, and $`\mathrm{\Psi }=(\psi _0,\psi _1,\mathrm{},\psi _n)`$, with $`\psi _i\stackrel{~}{๐ข}^{ni,i}`$. Then we define
$$\mathrm{\Psi },C_m=\underset{i=0}{\overset{n1}{}}\mathrm{exp}(\psi _i,c_i)\psi _n,c_n^{}.$$
(2.4.5)
By the very construction of the double complexes $`\stackrel{~}{๐ข}^,`$ and $`๐ฒ_,`$, the total differentials $`D`$ and $``$ are transpose to each other, namely
$$D\mathrm{\Psi },C_m=\mathrm{\Psi },C_m$$
for all $`\mathrm{\Psi }\stackrel{~}{๐ข}^,`$, $`C๐ฒ_,`$. The pairing (2.4.5) descends to the corresponding homology and cohomology groups and is non degenerate. It defines a pairing between $`H^{}(\mathrm{Tot}\stackrel{~}{๐ข}^,)`$ and $`H_{}(\mathrm{Tot}๐ฒ_,)`$ which we continue to denote by $`,_m`$.
It is easy to describe the relation between the multiplicative pairing $`,_m`$ and the $`/(p)`$-valued additive pairing introduced earlier. Namely, let $`\mathrm{\Phi }\mathrm{Tot}๐ข^,)`$, $`C๐ฒ_,`$ and let $`\mathrm{\Psi }`$ be the corresponding element in $`\stackrel{~}{๐ข}^,`$. The we have
$$\mathrm{\Psi },C_m=\mathrm{exp}\{\mathrm{\Phi },^{\mathrm{}}C/(2\pi i)^{p1}\}.$$
It what follows we will use freely both forms of the pairing, multiplicative and additive, depending on the context.
## 3 Construction of the action
### 3.1 General remarks
The next sections will be devoted to the detailed construction of the action functional โ or rather its exponential โ by specifying the following.
* A resolution of the Deligne complex $`(3)_๐^{}`$.
* A representative for a class in $`H_๐^3(X,(3))`$ that โstartsโ from a collection $`\{\omega _i[f]\}_{iI}`$ of โlocal Lagrangians densitiesโ โ top forms on $`X`$ โ defined with respect to a given covering $`๐ฐ_X=\{U_i\}_{iI}`$ of $`X`$.
The latter data come from Polyakovโs ansatz, with dynamical field given by a deformation map $`f:X\stackrel{~}{X}`$ and with external field given by a smooth projective connection of $`X`$. Before doing so, we make some remarks of general character.
* The Deligne complex $`(3)_๐^{}`$ is especially convenient for treating various logarithmic terms produced in descent calculations, while keeping additivity.
* The โlocal Lagrangianโ $`\mathrm{\Omega }[f]`$ appears as a total cocycle of total degree 3 in the Deligne complex $`(3)_๐^{}`$, and we define the action functional by evaluating this cocycle over the representative $`\mathrm{\Sigma }`$ of the fundamental class of the Riemann surface $`X`$
$$S[f]=\mathrm{\Omega }[f],\mathrm{\Sigma },$$
described in 2.4.2. According to 2.4.3, $`S[f]/(3)`$, so that the functional
$$A[f]=\mathrm{\Omega }[f],\mathrm{\Sigma }_m=\mathrm{exp}\{S[f]/(2\pi i)^2\}$$
is the *exponential* of the action.
* A similar approach was taken in in order to describe certain topological terms arising in two-dimensional quantum field theories. In our case the field is a deformation $`f:X\stackrel{~}{X}`$ and the procedure differs in that we construct the whole representing cocycle starting from one end of the descent staircase.
* According to the exponentials of action functionals should be more properly regarded as $`^{}`$-torsors rather than numbers. This is most apparent when dealing with manifolds with boundaries. A similar situation occurs in our case, when $`X`$ is a compact Riemann surface: the definition of the local Lagrangian cocycle $`\mathrm{\Omega }[f]`$ depends on the trivialization of the tame symbol $`(TX,TX]`$, described by an ($`f`$-independent) element of $`H^2(X,^{})^{}`$. As a result, the multiplicative action functional $`A[f]`$ is a $`^{}`$-torsor.
* The action functional $`A[f]`$, defined through hypercohomology admits the following geometric interpretation. According to Sect. 2.2, the group $`H_๐^3(X,(3))`$ classifies isomorphism classes of gerbes equipped with connective structure and curving . Since $`dimX=2`$, these are necessarily flat, therefore they are classified by their holonomy via the isomorphism $`H_๐^3(X,(3))H^2(X,^{})`$. Thus $`A[f]`$ can be interpreted as the holonomy of an appropriate higher algebraic structure.
### 3.2 Setup for regular ฤech coverings
Let $`๐ฐ_X=\{U_i\}_{iI}`$ be an open cover of $`X`$, which we assume to be a good cover, i.e. all nonempty intersections $`U_{i_0,\mathrm{},i_p}=U_{i_0}\mathrm{}U_{i_p}`$ are contractible. Therefore, we are in a ฤech-de Rham situation , and the double complex $`๐ข_๐^{p,q}\stackrel{def}{=}\stackrel{ห}{C}^q(๐ฐ_X,(3)_๐^p)`$ computes $`H_๐^{}(X,(3))`$. Let $`\{z_i:U_i\}_{iI}`$ be holomorphic coordinates for the complex structure of $`X`$, and let $`z_{ij}:z_j(U_iU_j)z_i(U_iU_j)`$ be coordinate change functions: $`z_i=z_{ij}z_j`$ on $`U_iU_j`$.
###### Remark 3.2.1.
One could also use coordinate functions with in $`^1`$ instead of $``$.
More generally, for $`U_{i_0,\mathrm{},i_q}`$ there are holomorphic coordinates $`z_{i_0},\mathrm{},z_{i_q}`$ with $`z_{i_k}=z_{i_ki_{k+1}}(z_{i_{k+1}})`$, $`k=0,\mathrm{},q1`$. If $`\varphi \stackrel{ห}{C}^{q1}(๐ฐ_X,\underset{ยฏ}{A}_X^p)`$ is a ฤech cochain, i.e. $`\varphi =\{\varphi _{i_0,\mathrm{},i_{q1}}\}`$, where the components $`\varphi _{i_0,\mathrm{},i_{q1}}`$ are $`p`$ forms on $`U_{i_0,\mathrm{},i_{q1}}`$ its ฤech differential is defined as
$$\stackrel{ห}{\delta }\varphi _{i_0,\mathrm{},i_q}=\underset{k=0}{\overset{q1}{}}(1)^k\varphi _{i_0,\mathrm{},\widehat{i_k},\mathrm{},i_q}+(1)^q(z_{i_{q1}i_q})^{}\varphi _{i_0,\mathrm{},i_{q1}}.$$
It is understood that each component $`\varphi _{i_0,\mathrm{},i_{q1}}`$ of a ฤech cochain $`\varphi `$ is expressed in the coordinate $`z_{i_{q1}}`$, i.e. the one determined by the last index, and we will use this convention throughout the paper.
Given a quasi-conformal map $`f:X\stackrel{~}{X}`$, denote by $`๐ฑ_X=\{V_i\}_{iI}`$, where $`V_i=f(U_i)`$, the corresponding good open cover for $`\stackrel{~}{X}`$. Let $`\{w_i:V_i\}_{iI}`$ be holomorphic coordinates for the complex structure of $`\stackrel{~}{X}`$, and let $`w_{ij}:w_j(V_iV_j)w_i(V_iV_j)`$ be the corresponding coordinate change functions: $`w_i=w_{ij}w_j`$ on $`V_iV_j`$. Let $`f_i=w_if|_{U_i}z_i^1,iI`$, be local representatives of the map $`f`$, satisfying the transformation law
$$f_iz_{ij}=w_{ij}f_j.$$
(3.2.1)
Denote $`f_i\stackrel{def}{=}f_i/z_i`$ and $`\overline{}_if_i\overline{}f_i\stackrel{def}{=}f_i/\overline{z}_i`$, and introduce local representatives of the Beltrami differential $`\mu `$ by $`\mu _i=\overline{}f_i/f_i`$.
It follows from (3.2.1) that
$$f_iz_{ij}z_{ij}^{}=w_{ij}^{}f_jf_j,$$
(3.2.2)
$$\overline{}f_iz_{ij}\overline{z_{ij}^{}}=w_{ij}^{}f_j\overline{}f_j,$$
(3.2.3)
and
$$\mu _iz_{ij}\frac{\overline{z_{ij}^{}}}{z_{ij}^{}}=\mu _j.$$
(3.2.4)
Since $`\xi _{ij}\stackrel{def}{=}z_{ij}^{}z_j=dz_i/dz_j`$ are transition functions for the holomorphic tangent bundle $`TX`$, and $`\stackrel{~}{\xi }_{ij}\stackrel{def}{=}w_{ij}^{}w_j=dw_i/dw_j`$ are the transition functions for $`T\stackrel{~}{X}`$, it follows from (3.2.2) that $`f`$ is a section of the bundle $`T^{}Xf^1T\stackrel{~}{X}`$, or $`fA^{1,0}(f^1T\stackrel{~}{X})`$. Here $`f^1T\stackrel{~}{X}`$ is the pull-back of the tangent bundle over $`\stackrel{~}{X}`$ by $`f`$. Similarly, $`\overline{}fA^{0,1}(f^1T\stackrel{~}{X})`$.
In the $`C^{\mathrm{}}`$ category $`f^1T\stackrel{~}{X}TX`$, so that $`T^{}Xf^1T\stackrel{~}{X}`$ is isomorphic to the trivial bundle. This is also implied directly by the transition formula (3.2.2), since $`f_i0`$, $`f`$ being a diffeomorphism. Thus $`f`$ is an explicit trivializing section for $`T^{}Xf^1T\stackrel{~}{X}`$, that establishes the isomorphism between $`T^{}Xf^1T\stackrel{~}{X}`$ and the trivial line bundle.
Introducing representatives $`c_{ijk}`$ and $`\stackrel{~}{c}_{ijk}`$ for the first Chern classes $`c_1(TX)=c_1(\stackrel{~}{T}X)`$, we have
$`c_{ijk}`$ $`=\stackrel{ห}{\delta }(\{\mathrm{log}z_{}^{}\})_{ijk},`$ (3.2.5a)
$`\stackrel{~}{c}_{ijk}`$ $`=\stackrel{ห}{\delta }(\{\mathrm{log}w_{}^{}f_{}\})_{ijk},`$ (3.2.5b)
$`b_{ij}`$ $`=\mathrm{log}w_{ij}^{}f_j\mathrm{log}z_{ij}^{}\mathrm{log}f_iz_{ij}+\mathrm{log}f_j,`$ (3.2.5c)
and, obviously, $`\stackrel{ห}{\delta }(\{b_{}\})_{ijk}=\stackrel{~}{c}_{ijk}c_{ijk}`$. All the numbers $`b_{ij}`$, $`c_{ijk}`$ and $`\stackrel{~}{c}_{ijk}`$ are in $`(1)`$.
Although one can get $`\stackrel{~}{c}_{ijk}=c_{ijk}`$ and $`b_{ij}=0`$ through a suitable redefinition of the logarithm branches, there is no additional complication (except, perhaps, the notation) in keeping the general situation.
### 3.3 The local Lagrangian cocycle
In order to construct the action functional, one needs an ansatz for its top degree part. Following , we promote the standard Polyakovโs chiral action<sup>6</sup><sup>6</sup>6More precisely, Polyakovโs chiral action has no second term in (3.3.1), which, in fact, is not necessary in genus zero.,
$$\omega _i=\frac{^2f_i}{f_i}\mu _idz_id\overline{z}_i+2\mu _ih_idz_id\overline{z}_i,$$
(3.3.1)
to an element $`\{2\pi \sqrt{1}\omega _i\}_{iI}๐ข_๐^{3,0}`$. Here $`h=\{h_i\}_{iI}`$ is a $`C^{\mathrm{}}`$ coboundary for the Schwarzian cocycle
$$\{z_i,z_j\}=\frac{d^3z_i}{dz_j^3}\frac{3}{2}\left(\frac{d^2z_i}{dz_j^2}\right)^2,$$
relative to the cover $`๐ฐ_X`$ (see ). In other words, it satisfies the following transformation law
$$\{z_i,z_j\}=h_jh_iz_{ij}(z_{ij}^{})^2$$
(3.3.2)
on $`U_iU_j`$. Clearly, such an $`h`$ exists, since the Schwarzian cocycle is already zero in the holomorphic category . The space $`๐ฌ(X)`$ of all such $`h`$ includes the holomorphic projective connections, and is an affine space over the vector space $`H^0(X,(\underset{ยฏ}{A}_X^{1,0})^2)`$. Let us call such an $`h`$ a *smooth* projective connection (even though that we do not relate it to projective structures).
Following the usual strategy of descending the staircase in the double complex $`๐ข_๐^,`$, starting with the 0-cochain $`\{\omega _i\}`$ of 2-forms on $`X`$, we find a 1-cochain of 1-forms $`\{\theta _{ij}\}`$ and a 2-cochain of functions $`\{\mathrm{\Theta }_{ijk}\}`$ satisfying
$`\stackrel{ห}{\delta }(\omega _{})_{ij}`$ $`=d\theta _{ij}`$
$`\stackrel{ห}{\delta }(\theta _{})_{ijk}`$ $`=d\mathrm{\Theta }_{ijk}.`$
Imposing the condition $`\stackrel{ห}{\delta }\mathrm{\Theta }=0mod(2)`$ ensures that the total element
$$\mathrm{\Omega }\stackrel{def}{=}2\pi \sqrt{1}(\{\omega _i\},\{\theta _{ij}\},\{\mathrm{\Theta }_{ijk}\},\{m_{ijkl}\})$$
where $`m=\stackrel{ห}{\delta }\mathrm{\Theta }`$, is a cocycle in the total complex.
Solvability of the descent equations is proved in the standard way using the acyclic property of the good cover $`๐ฐ_X`$ and Poincarรฉ lemma on differential forms. Namely, $`\stackrel{ห}{\delta }d\omega =0`$ implies $`\stackrel{ห}{\delta }\omega =d\theta `$ and $`0=\stackrel{ห}{\delta }d\theta =d\stackrel{ห}{\delta }\theta `$ implies $`\stackrel{ห}{\delta }\theta =d\mathrm{\Theta }`$. Finally, from $`\stackrel{ห}{\delta }d\mathrm{\Theta }=d\stackrel{ห}{\delta }\mathrm{\Theta }=0`$ one concludes $`\stackrel{ห}{\delta }\mathrm{\Theta }\stackrel{ห}{Z}^3(๐ฐ_X,_X)`$. From de Rham theorem $`\stackrel{ห}{H}^p(X,)H_{dR}^p(X)`$ it follows for dimensional reasons that $`\stackrel{ห}{\delta }\mathrm{\Theta }=0`$, after possible rescaling of constants.
The foregoing shows that one can get a โminimalโ cocycle with the condition $`m_{ijkl}=0`$, albeit not in explicit form. However, our goal is to have a cocycle $`\mathrm{\Omega }[f]`$ with โgoodโ dependence on the dynamical field $`f`$ (i.e. with the same variational properties as in the genus zero case). It is most remarkable that such cocycle $`\mathrm{\Omega }[f]`$ can in fact be computed explicitly, allowing for a geometric interpretation as to why $`\stackrel{ห}{\delta }\mathrm{\Theta }=0mod(2)`$. This computation is accomplished in the following steps.
#### 3.3.1 $`\stackrel{ห}{\delta }\omega =d\theta `$:
We find, using the transformation rules (3.2.2)- (3.2.4),
$$\begin{array}{cc}\hfill \stackrel{ห}{\delta }\omega _{ij}& =\omega _j\omega _i\hfill \\ & =\begin{array}{cc}& d\left(2\mu _j\frac{z_{ij}^{\prime \prime }}{z_{ij}^{}}d\overline{z}_j\left(\mathrm{log}(w_{ij}^{}f_j)+\mathrm{log}z_{ij}^{}\right)d\mathrm{log}f_j+\mathrm{log}(w_{ij}^{}f_j)d\mathrm{log}z_{ij}^{}\right)\hfill \\ & +2\mu _jh_jdz_jd\overline{z}_j2\mu _ih_idz_id\overline{z}_i2\mu _j\{z_i,z_j\}dz_jd\overline{z}_j.\hfill \end{array}\hfill \end{array}$$
(3.3.3)
In light of (3.3.2), equation (3.3.3) reads
$$\stackrel{ห}{\delta }\omega =d\theta ,$$
with $`\theta `$ given by the first two terms on the RHS of (3.3.3), that is,
$$\begin{array}{cc}\hfill \theta _{ij}=2\mu _j\frac{z_{ij}^{\prime \prime }}{z_{ij}^{}}d\overline{z}_j& \left(\mathrm{log}(w_{ij}^{}f_j)+\mathrm{log}z_{ij}^{}\right)d\mathrm{log}f_j\hfill \\ & +\mathrm{log}(w_{ij}^{}f_j)d\mathrm{log}z_{ij}^{}.\hfill \end{array}$$
(3.3.4)
#### 3.3.2 $`\stackrel{ห}{\delta }\theta `$:
The first term on the RHS of (3.3.4) is a cocycle, as it has the ฤech cup product of two terms which are cocycles themselves. We can ignore it from now on. The term on the second line of (3.3.4) is also cup product, so its coboundary is computed by applying $`\stackrel{ห}{\delta }(ab)=\stackrel{ห}{\delta }(a)b+(1)^{\mathrm{deg}a}a\stackrel{ห}{\delta }(b)`$. For the remaining term the cocycle is computed directly. The final result is
$$\begin{array}{cc}\hfill \stackrel{ห}{\delta }\left(\theta \right)_{ijk}=& \mathrm{log}w_{ij}^{}d\mathrm{log}w_{jk}^{}+\mathrm{log}z_{ij}^{}d\mathrm{log}z_{jk}^{}\hfill \\ & (\stackrel{~}{c}_{ijk}+c_{ijk})d\mathrm{log}f_kd\left(\mathrm{log}z_{ij}^{}\mathrm{log}w_{jk}^{}\right)\hfill \\ & +\stackrel{~}{c}_{ijk}d\mathrm{log}z_{ik}^{},\hfill \end{array}$$
(3.3.5)
where we suppressed the $`f`$-dependence. To restore it, notice that on the triple intersection $`U_iU_jU_k`$ everything is evaluated with respect to the coordinate $`z_k`$, so that $`\mathrm{log}w_{ij}^{}f_j|_{U_k}=\mathrm{log}w_{ij}^{}f_jz_{jk}\mathrm{log}w_{ij}^{}w_{jk}f_k`$. We shall use this convention in the sequel, in order to keep some of the expressions less cumbersome.
#### 3.3.3 $`\stackrel{ห}{\delta }\theta =d\mathrm{\Theta }`$:
Here we are using Deligne tame symbols in holomorphic category, introduced in 2.2 in order to find $`\mathrm{\Theta }`$ satisfying the equation $`\stackrel{ห}{\delta }\theta =d\mathrm{\Theta }`$ and to check that $`\stackrel{ห}{\delta }\mathrm{\Theta }=0mod(2)`$.
Consider the tame symbol $`(TX,TX]`$, which is represented in ฤech cohomology by the element
$$(\mathrm{log}z_{ij}^{}d\mathrm{log}z_{jk}^{},c_{ijk}\mathrm{log}z_{kl}^{},c_{ijk}c_{klm})\stackrel{ห}{C}^2(\underset{ยฏ}{\mathrm{\Omega }}_X^1)\stackrel{ห}{C}^3(\underset{ยฏ}{๐ช}_X)\stackrel{ห}{C}^4((2)_X),$$
where $`\{c_{ijk}\}`$ represents the first Chern class of $`TX`$. As we mentioned in section 2.2,
$$^4(X,(2)_{๐,\mathrm{โ๐๐}}^{})H^3(X,^{})=0,$$
so that the total cocycle representing $`(TX,TX]`$ is a coboundary:
$$(\mathrm{log}z_{ij}^{}d\mathrm{log}z_{jk}^{},c_{ijk}\mathrm{log}z_{kl}^{},c_{ijk}c_{klm})=D(\tau _{ij},\varphi _{ijk},n_{ijkl}),$$
where $`(\tau _{ij})\stackrel{ห}{C}^1(๐ฐ_X,\underset{ยฏ}{\mathrm{\Omega }}_X^1)`$, $`(\varphi _{ijk})\stackrel{ห}{C}^2(๐ฐ_X,\underset{ยฏ}{๐ช}_X)`$ and $`(n_{ijkl})\stackrel{ห}{C}^3(๐ฐ_X,(2)_X)`$. Computing the RHS yields the relations
$`\mathrm{log}z_{ij}^{}d\mathrm{log}z_{jk}^{}`$ $`=(\stackrel{ห}{\delta }\tau )_{ijk}+d\varphi _{ijk}`$ (3.3.6a)
$`c_{ijk}\mathrm{log}z_{kl}^{}`$ $`=(\stackrel{ห}{\delta }\varphi )_{ijkl}+n_{ijkl}`$ (3.3.6b)
$`c_{ijk}c_{klm}`$ $`=(\stackrel{ห}{\delta }n)_{ijklm}.`$ (3.3.6c)
There is an entirely similar situation for the deformed Riemann Surface $`\stackrel{~}{X}`$ and the corresponding symbol $`(T\stackrel{~}{X},T\stackrel{~}{X}]`$, for which we introduce the corresponding objects $`\stackrel{~}{\tau }_{ij}`$, $`\stackrel{~}{\varphi }_{ijk}`$ and $`\stackrel{~}{n}_{ijkl}`$. Using these results we rewrite $`\stackrel{ห}{\delta }\theta `$ as
$$\begin{array}{cc}\hfill \stackrel{ห}{\delta }\left(\theta \right)_{ijk}& =\stackrel{ห}{\delta }(f^{}(\stackrel{~}{\tau }))_{ijk}+df^{}(\stackrel{~}{\varphi }_{ijk})\stackrel{ห}{\delta }(\tau )_{ijk}d\varphi _{ijk}\hfill \\ & (\stackrel{~}{c}_{ijk}+c_{ijk})d\mathrm{log}f_kd\left(\mathrm{log}z_{ij}^{}\mathrm{log}w_{jk}^{}\right)\hfill \\ & +\stackrel{~}{c}_{ijk}d\mathrm{log}z_{ik}^{},\hfill \end{array}$$
where $`f^{}(\stackrel{~}{\tau }_{ij})`$ and $`f^{}(\stackrel{~}{\varphi }_{ijk})`$ are pull-backs of forms $`\stackrel{~}{\tau }_{ij}`$ and $`\stackrel{~}{\varphi }_{ijk}`$ on $`X`$. Now, perform the shift:
$$\theta _{ij}\widehat{\theta }_{ij}\stackrel{def}{=}\theta _{ij}f^{}(\stackrel{~}{\tau }_{ij})+\tau _{ij}.$$
This is possible since $`\tau _{ij}`$ and $`\stackrel{~}{\tau }_{ij}`$ are holomorphic relative to the respective complex structures, implying $`d\tau _{ij}=0`$ and $`df^{}(\stackrel{~}{\tau }_{ij})=0`$, so that
$$d\widehat{\theta }_{ij}=d\theta _{ij}=\stackrel{ห}{\delta }(\omega )_{ij},$$
without affecting the 2-form part of the action.
From now on we assume that $`\theta _{ij}`$ has been redefined in this way, that is
$$\theta _{ij}=\theta _{ij}^{\mathrm{๐๐๐}}f^{}(\stackrel{~}{\tau }_{ij})+\tau _{ij},$$
(3.3.7)
where $`\theta _{ij}^{\mathrm{๐๐๐}}`$ is given by formula (3.3.4), and we can finally put
$$\stackrel{ห}{\delta }\theta =d\mathrm{\Theta },$$
with
$$\begin{array}{cc}\hfill \mathrm{\Theta }_{ijk}& =f^{}(\stackrel{~}{\varphi }_{ijk})\varphi _{ijk}(\stackrel{~}{c}_{ijk}+c_{ijk})\mathrm{log}f_k\hfill \\ & \mathrm{log}z_{ij}^{}\mathrm{log}w_{jk}^{}+\stackrel{~}{c}_{ijk}\mathrm{log}z_{ik}^{}.\hfill \end{array}$$
(3.3.8)
#### 3.3.4 $`\stackrel{ห}{\delta }\mathrm{\Theta }`$:
Using the relations (3.3.6) we compute:
$$\stackrel{ห}{\delta }\mathrm{\Theta }_{ijkl}=\stackrel{~}{n}_{ijkl}n_{ijkl}(\stackrel{~}{c}_{ijk}+c_{ijk})b_{kl}+c_{ijl}\stackrel{~}{c}_{jkl}c_{ikl}\stackrel{~}{c}_{ijk},$$
so that $`\stackrel{ห}{\delta }\mathrm{\Theta }\stackrel{ห}{C}^3(๐ฐ_X,(2)_X)`$. Setting
$$m_{ijkl}\stackrel{def}{=}(\stackrel{ห}{\delta }\mathrm{\Theta })_{ijkl},$$
we can summarize the foregoing in the following
###### Proposition 3.3.1.
The total cochain
$$\mathrm{\Omega }\stackrel{def}{=}2\pi \sqrt{1}(\omega _i,\theta _{ij},\mathrm{\Theta }_{ijk},m_{ijkl}),$$
with $`\omega _i`$ given by the Polyakov form (3.3.1), represents a class in $`H_๐^3(X,(3))`$.
###### Proof.
All the preceding computations amount to show that
$$D\mathrm{\Omega }=2\pi \sqrt{1}(\omega _j+\omega _i+d\theta _{ij},(\stackrel{ห}{\delta }\theta )_{ijk}d\mathrm{\Theta }_{ijk},(\stackrel{ห}{\delta }\mathrm{\Theta })_{ijkl}m_{ijkl},(\stackrel{ห}{\delta }m)_{ijklp})=0.$$
Then $`\mathrm{\Omega }`$ represents a class since the double complex $`๐ข_๐^,`$ computes the hypercohomology. โ
Now that we have constructed the Lagrangian cocycle from the Polyakov top form in (3.3.1), we can finally give the
###### Definition 3.3.2.
Let $`\mu (X)`$ be a Beltrami coefficient, $`f`$ be the associated deformation map, and $`\mathrm{\Omega }[f]`$ be the local Lagrangian cocycle constructed from (3.3.1). The *Polyakov action* functional on $`X`$ is given by the evaluation
$$S[f]\stackrel{def}{=}\mathrm{\Omega }[f],\mathrm{\Sigma },$$
(3.3.9)
over the representative $`\mathrm{\Sigma }`$ of the fundamental class of $`X`$ given in 2.4.2 and in the appendix.
###### Remark 3.3.3.
As it follows from the definition, Polyakovโs action is well-defined modulo $`(3)`$, so that only its exponential $`A[f]=\mathrm{exp}\{S[f]/(2\pi i)^2\}`$ is well-defined. It also follows from the definition of the pairing in section 2.4.2 that the functional $`A[f]`$ actually depends only on the cohomology class in $`H_๐^3(X,(3))`$ of the local Lagrangian cocycle $`\mathrm{\Omega }[f]`$.
By construction, the cocycle $`\mathrm{\Omega }[f]`$ depends also on a smooth projective connection $`h๐ฌ(X)`$, so that the exponential of the action defines the map $`A:๐ฌ(X)\times (X)^{}`$, where the dependence on the first factor is that of an external field.
Here we analyze the dependence of the action functionals $`S[f]`$ and $`A[f]`$ on the choice of the logarithm branches. We also study the trivializing coboundary for the tame symbol $`(TX,TX]`$, analyze the dependence of the action on this trivialization, and show that $`A[f]`$ should be in fact considered as taking its values in a $`^{}`$-torsor.
#### 3.3.5 Dependency on logs
Here we prove the following
###### Proposition 3.3.4.
The functional $`A[f]`$ is independent of the choice of the logarithm branches in (3.2.5).
###### Proof.
It is sufficient to show that changing the definition of the various logarithm branches in $`\mathrm{\Omega }`$ amounts to change it by a coboundary. First, we change these branches,
$$\mathrm{log}z_{ij}^{}\mathrm{log}z_{ij}^{}+k_{ij}$$
$$\mathrm{log}w_{ij}^{}\mathrm{log}w_{ij}^{}+\stackrel{~}{k}_{ij}$$
$$\mathrm{log}f_i\mathrm{log}f_i+p_i$$
where $`k_{ij},\stackrel{~}{k}_{ij},p_i(1)`$. The effect of these changes on the representatives of the Chern classes of $`TX`$ and $`T\stackrel{~}{X}`$ is
$$b_{ij}b_{ij}+\stackrel{~}{k}_{ij}k_{ij}+p_jp_i$$
$$c_{ijk}c_{ijk}+\stackrel{ห}{\delta }(k)_{ijk}$$
$$\stackrel{~}{c}_{ijk}\stackrel{~}{c}_{ijk}+\stackrel{ห}{\delta }(\stackrel{~}{k})_{ijk}.$$
While the term $`\omega _i`$ is obviously invariant under these changes, $`\theta _{ij}`$ and $`\mathrm{\Theta }_{ijk}`$, by descent theory, transform as follows
$$\theta _{ij}\theta _{ij}+d\psi _{ij},$$
$$\mathrm{\Theta }\mathrm{\Theta }+\stackrel{ห}{\delta }\psi r_{ijk},$$
where $`\psi \stackrel{ห}{C}^1(๐ฐ_X,\underset{ยฏ}{A}_X^0)`$ and $`r\stackrel{ห}{C}^2(๐ฐ_X,)`$. Note that if $`r_{ijk}(2)`$ for any $`ijk`$, then $`\mathrm{\Omega }\mathrm{\Omega }+D\lambda `$, where $`\lambda =(0,\psi _{ij},r_{ijk})`$, and we are done.
To prove that $`r\stackrel{ห}{C}^2(๐ฐ_X,(2))`$, we actually compute the shift for $`\mathrm{\Theta }`$. First, we explicitly determine
$$\psi _{ij}=(\stackrel{~}{k}_{ij}+k_{ij})\mathrm{log}f_j+\stackrel{~}{k}_{ij}\mathrm{log}z_{ij}^{}.$$
Next, we explicitly compute the shift of the total cocycle representing $`(TX,TX]`$. This is a straightforward calculation, using relations (3.3.6), with the result:
$`\tau _{ij}`$ $`\tau _{ij}`$
$`\varphi _{ijk}`$ $`\varphi _{ijk}k_{ij}\mathrm{log}z_{jk}^{}`$
$`n_{ijkl}`$ $`n_{ijkl}+k_{ij}c_{jkl}+c_{ijk}k_{kl}+(\stackrel{ห}{\delta }k)_{ijk}k_{kl}.`$
Similar formulas are valid for the shift of $`(T\stackrel{~}{X},T\stackrel{~}{X}]`$. Putting everything together, we get
$$\begin{array}{c}r_{ijk}=(\stackrel{~}{k}_{ij}+k_{ij})b_{jk}+\left(\stackrel{~}{c}_{ijk}+c_{ijk}+(\stackrel{ห}{\delta }k)_{ijk}+(\stackrel{ห}{\delta }\stackrel{~}{k})_{ijk}\right)p_k\hfill \\ \hfill +k_{ij}\stackrel{~}{k}_{jk}\stackrel{~}{c}_{ijk}k_{ik}(\stackrel{ห}{\delta }\stackrel{~}{k})_{ijk}k_{ik}+\stackrel{~}{k}_{jk}c_{ijk}+\stackrel{~}{k}_{ij}c_{ijk}(2).\end{array}$$
(3.3.10)
#### 3.3.6 A more detailed analysis of the vanishing tame symbol
Here we analyze the condition $`(TX,TX]=0`$ as an element of $`^4(X,(2)_{๐,\mathrm{โ๐๐}}^{})`$ in more detail. In particular, we investigate the possibility of putting the trivializing cochains $`(\tau _{ij},\varphi _{ijk},n_{ijkl})`$ and $`(\stackrel{~}{\tau }_{ij},\stackrel{~}{\varphi }_{ijk},\stackrel{~}{n}_{ijkl})`$ into some specific forms. This analysis is based on the relations (3.3.6), which we rewrite here:
$`\mathrm{log}z_{ij}^{}d\mathrm{log}z_{jk}^{}`$ $`=(\stackrel{ห}{\delta }\tau )_{ijk}+d\varphi _{ijk}`$
$`c_{ijk}\mathrm{log}z_{kl}^{}`$ $`=(\stackrel{ห}{\delta }\varphi )_{ijkl}+n_{ijkl}`$
$`c_{ijk}c_{klm}`$ $`=(\stackrel{ห}{\delta }n)_{ijklm}.`$
The first equation above calls for the differential equation
$$\mathrm{log}z_{ij}^{}z_{jk}d\mathrm{log}z_{jk}^{}=dL_{ijk}.$$
Its solution $`L_{ijk}(z_k)`$ can be considered as a Bloch dilogarithm associated to the symbol $`(z_{ij}^{},z_{jk}^{}]`$, which is the cup-product in Deligne cohomology of the two invertible functions $`z_{ij}^{}`$ and $`z_{jk}^{}`$ and is a trivial element of $`H_๐^2(U_{ijk},(2))`$ (see for more details). The consistency condition on quadruple intersections $`U_{ijkl}`$ is obtained by applying the ฤech coboundary to the differential equation satisfied by $`L_{ijk}`$. One gets
$$c_{ijk}\mathrm{log}z_{kl}^{}=(\stackrel{ห}{\delta }L)_{ijkl}+\alpha _{ijkl},$$
where $`\alpha _{ijkl}`$ is a $``$-valued cochain โ an integration constant. By taking the ฤech coboundary of the last relation we get
$$c_{ijk}c_{klm}=(\stackrel{ห}{\delta }\alpha )_{ijklm}.$$
Therefore,
$$\stackrel{ห}{\delta }(\alpha n)=0,$$
that is, the element $`\alpha n`$ is a 3-cocycle. By dimensional reasons, it must be a coboundary,
$$\alpha =n+\stackrel{ห}{\delta }\beta ,$$
with $`\beta `$ being a $`2`$-cochain with values in $``$. It follows that
$$c_{ijk}\mathrm{log}z_{kl}^{}=\stackrel{ห}{\delta }(L\beta )_{ijkl}+n_{ijkl}.$$
As a result, we effectively obtained a trivializing cocycle for the tame symbol $`(TX,TX]`$ which does not include a 1-form:
$$(\mathrm{log}z_{ij}^{}d\mathrm{log}z_{jk}^{},c_{ijk}\mathrm{log}z_{kl}^{},c_{ijk}c_{klm})=D(0,L_{ijk},n_{ijkl}),$$
where we relabeled $`L\beta L`$.
#### 3.3.7 Relation with $`^{}`$-torsors
Notice that the trivialization of the tame symbol $`(TX,TX]`$ is defined up to a cocycle representing an element in $`^3(X,(2)_{๐,\mathrm{โ๐๐}}^{})H^2(X,/(2))H^2(X,^{})^{}`$. Thus there is a $`^{}`$-action on the functional $`A[f]`$ which simply is the shift of the total trivializing cochain $`(\tau _{ij},\varphi _{ijk},n_{ijkl})`$ by a cocycle representing a class in $`^3(X,(2)_{๐,\mathrm{โ๐๐}}^{})`$. From this it is clear that, keeping $`f`$ fixed, the functional $`A[f]`$ does not simply take its values in $`^{}`$, but rather in a $`^{}`$-torsor $`T`$. From this perspective, choosing a specific total cochain to trivialize the symbol $`(TX,TX]`$ amounts to choosing an isomorphism $`T\stackrel{}{}`$.
The $`^{}`$-action can be described explicitly if we make use of the cocycle $`(0,L_{ijk},n_{ijkl})`$, obtained by choosing a dilogarithm $`L_{ijk}`$ for the symbol $`(z_{ij}^{},z_{jk}^{}]`$. Namely, as it follows from the discussion in the previous section, we can add to $`L_{ijk}`$ a cocycle $`(\beta _{ijk},p_{ijkl})`$ representing an element in
$$^3(X,(2)\stackrel{\mathit{ฤฑ}}{})H^2(X,^{}).$$
Note that, by definition, $`\stackrel{ห}{\delta }\beta =p(2)`$.
Since the action functional is defined using trivialization of *two* tame symbols, $`(TX,TX]`$ and $`(T\stackrel{~}{X},T\stackrel{~}{X}]`$, the above argument should be applied to both cochains $`(\tau _{ij},\varphi _{ijk},n_{ijkl})`$ and $`(\stackrel{~}{\tau }_{ij},\stackrel{~}{\varphi }_{ijk},\stackrel{~}{n}_{ijkl})`$, so that we have in fact two $`^{}`$-actions. From a Teichmรผller theory point of view, these two actions refer to very different structures. One is defined in terms of the complex structure $`X`$ which is fixed throughout (a base point in Teichmรผller space), while the other is relative to the $`f`$-dependent complex structure $`\stackrel{~}{X}`$. The latter action depends on the dynamical field $`f`$.
Thus it is appropriate to speak of a $`(^{},^{})`$-action, in the sense that the space $`T`$ where the action takes its values carries two simultaneous (and compatible) $`^{}`$-actions.
### 3.4 Other coverings โ a dictionary
In this section we set up a dictionary connecting generalized ฤech formalism developed in 2.3 and 2.4 with the formalism used in for the universal cover of $`X`$. Besides comparing the two formalisms, by applying the dictionary to the formulas in 3.3, we also clarify the explicit form of the Lagrangian cocycle constructed in . Specifically, we treat the โintegration constantsโ arising from solving the descent equations via Deligne complexes and analyze explicit dependence of the action functional on background projective structures.
#### 3.4.1
Start from the universal cover $`UX`$, which we specify as the upper half-plane $``$. Then $`\mathrm{Deck}(/X)\pi _1(X)\mathrm{\Gamma }`$, a finitely-generated, purely hyperbolic Fuchsian group (a discrete subgroup of $`\mathrm{PSL}_2()`$), uniformizing the Riemann surface $`X`$. The group $`\mathrm{\Gamma }`$ acts on $``$ by Mรถbius transformations.
Geometric objects on $`X`$ correspond to $`\mathrm{\Gamma }`$-equivariant objects on $``$: a tensor $`\varphi A^{p,q}(X)`$ corresponds to an automorphic form $`\varphi `$ for $`\mathrm{\Gamma }`$ of weight $`(2p,2q)`$, i.e. a function (indicated by the same name) $`\varphi :`$ such that
$$\varphi \gamma (\gamma ^{})^p(\overline{\gamma ^{}})^q=\varphi ,\gamma \mathrm{\Gamma }.$$
Clearly, an automorphic form is just a zero cocycle on $`\mathrm{\Gamma }`$ with values in $`\underset{ยฏ}{A}^{p,q}()`$. Examples of automorphic forms of geometric origin are provided by Beltrami differentials on $`X`$, that correspond to forms of weight $`(2,2)`$, by abelian differentials on $`X`$ โ global sections of $`\underset{ยฏ}{\mathrm{\Omega }}_X`$ โ that correspond to holomorphic forms of weight $`(2,0)`$, and by quadratic differentials on $`X`$ โ global sections of $`\underset{ยฏ}{\mathrm{\Omega }}_X^2`$ โ that correspond to holomorphic forms of weight $`(4,0)`$.
The deformation map $`f`$ is realized as a quasi-conformal map
$$f:๐ป$$
satisfying on $``$ the Beltrami equation
$$f_{\overline{z}}=\mu f_z,$$
(3.4.1)
where $`\mu `$ is a Beltrami differential for $`\mathrm{\Gamma }`$ on $``$ such that $`\mu _{\mathrm{}}<1`$. The Beltrami equation on $``$ should be supplemented by boundary conditions, that guarantee the following.
1. $`๐ป=f()`$ is a *quasi-disk*, i.e. a domain in $`^1`$ bounded by a closed Jordan curve and analytically isomorphic to $``$;
2. $`\stackrel{~}{\mathrm{\Gamma }}=f\mathrm{\Gamma }f^1\mathrm{PSL}_2()`$ is a discrete subgroup, isomorphic to $`\mathrm{\Gamma }`$ as an abstract group, acting on $`๐ป`$, i.e. a so-called *quasi-fuchsian* group. The isomorphism $`\mathrm{\Gamma }\stackrel{~}{\mathrm{\Gamma }}`$ intertwines $`f`$.
These boundary conditions are specified by extending $`\mu `$ to the whole complex plane, where the Beltrami equation has a unique solution up to a post-composition with Mรถbius transformation . The following two types are of particular importance.
1. Extension of $`\mu `$ by reflection to the lower half plane $`\overline{}`$: $`\mu (z)\stackrel{def}{=}\overline{\mu (\overline{z})}`$ for $`z\overline{}`$. Then $`๐ป=`$ and $`\stackrel{~}{\mathrm{\Gamma }}`$ is also a Fuchsian group.
2. Extension of $`\mu `$ by setting $`\mu (z)=0`$ for $`z\overline{}`$. In this case $`๐ป`$ is a quasi-disc and the dependence of the mapping $`f`$ on $`\mu `$ is holomorphic.
The formalism developed below will be independent of a particular boundary condition chosen.
#### 3.4.2
Here we address a minor normalization problem caused by the fact that the action of $`\mathrm{PSL}_2()`$ โ and therefore of $`\mathrm{\Gamma }`$ and $`\stackrel{~}{\mathrm{\Gamma }}`$ โ by Mรถbius transformations is on the left instead of on the right, as we assumed in 2.3. Assuming a right action yields all the standard formulas in group cohomology. On the other hand, a left action of $`\mathrm{\Gamma }`$ is more convenient in view of the fact that $``$ itself is the quotient of a principal fibration: $`\mathrm{PSL}_2()/\mathrm{SO}(2)`$.<sup>7</sup><sup>7</sup>7In general, we prefer to consider *right* principal fibrations. As a result, the surface itself is presented as a double coset space: $`X\mathrm{\Gamma }\backslash \mathrm{PSL}_2()/\mathrm{SO}(2)`$.
For a left action $`G\times UU`$ for a $`G`$-space $`UM`$ there is the isomorphism
$$\underset{q+1}{\underset{}{U\times _M\mathrm{}\times _MU}}G^q\times U,$$
sending the $`q`$-tuple $`(x_0,\mathrm{},x_q)`$ to the tuple $`(g_1,\mathrm{},g_q,x)`$ such that
$$(x_0,\mathrm{},x_q)=(g_1\mathrm{}g_qx,g_2\mathrm{}g_qx,\mathrm{},g_qx,x).$$
This arrangement makes the face maps $`d_i`$ appear in backward order, that is
$$d_0(g_1,\mathrm{},g_q,x)=(g_2,\mathrm{},g_q,x)\mathrm{}d_q(g_1,\mathrm{},g_q,x)=(g_1,\mathrm{},g_{q1},g_qx).$$
As a result, the action on the coefficients would be on the right and the coboundary operator $`\delta `$ in group cohomology should actually be read from right to left, as in
$$\begin{array}{cc}\hfill (\delta \varphi )_{g_1,\mathrm{},g_q}=\varphi _{g_2,\mathrm{},g_q}& +\underset{i=1}{\overset{q1}{}}(1)^i\varphi _{g_1,\mathrm{},g_ig_{i+1},\mathrm{},g_q}\hfill \\ & +(1)^qg_{q}^{}{}_{}{}^{}\varphi _{g_1,\mathrm{},g_{q1}},\hfill \end{array}$$
(3.4.2)
for $`\varphi `$ a $`(q1)`$ cochain. Observe that the pull-back action on the coefficients is a right one.
The familiar formulas in group cohomology can be retrieved by turning the left action into a right one using the standard trick
$$xg\stackrel{def}{=}g^1x,gG,xU,$$
which at the level of nerves amounts to performing the swap $`(g_1,\mathrm{},g_q,x)(x,g_q^1,\mathrm{},g_1^1)`$ in degree $`q`$. It follows that one has to evaluate all cochains over inverses of group elements. This is the convention we followed in .
On the other hand, given the action of $`\mathrm{\Gamma }`$ on $``$ as a left one, keeping the non standard form (3.4.2) parallels more closely the ฤech framework if we consider the pair $`(\gamma (z),z)\times _X`$, for $`z`$ and $`\gamma \mathrm{\Gamma }`$, as a change of coordinates, much like the pair $`(z_i,z_j)U_i\times _XU_jU_iU_j`$ with $`z_i=z_{ij}(z_j)`$. More generally, for $`U_{i_0}\mathrm{}U_{i_q}`$ there are coordinates $`z_{i_0},\mathrm{},z_{i_q}`$ with $`z_{i_k}=z_{i_ki_{k+1}}(z_{i_{k+1}})`$, $`k=0,\mathrm{},q1`$, and if $`\varphi \stackrel{ห}{C}^{q1}(๐ฐ_X,\underset{ยฏ}{A}^p)`$ then we have
$$\stackrel{ห}{\delta }\varphi _{i_0,\mathrm{},i_q}=\underset{k=0}{\overset{q1}{}}(1)^k\varphi _{i_0,\mathrm{},\widehat{i_k},\mathrm{},i_q}+(1)^q(z_{i_{q1}i_q})^{}\varphi _{i_0,\mathrm{},i_{q1}},$$
(3.4.3)
where the convention is that each component is expressed in the coordinate determined by the last index. This is the formula we used when performing explicit computations with ฤech cochains for the calculation of the local Lagrangian cocycle. Thus (3.4.3) becomes formally equal to (3.4.2) when we interpret the last pull-back by $`g_q`$ as the restriction isomorphism expressing everything in terms of the last coordinate.
#### 3.4.3
The translation of the constructions in 3.2 and 3.3 to the upper-half plane is now done according to the following table:
| ฤech:$`๐ฐ_X`$ | Upper-half plane $``$ |
| --- | --- |
| $`U_{i_0}\mathrm{}U_{i_n}`$ | $`\mathrm{\Gamma }^n\times `$ |
| $`z_{i_0},\mathrm{},z_{i_n}`$ | $`\gamma _1,\mathrm{},\gamma _n,z`$ |
| $`z_{i_k}=z_{i_ki_{k+1}}(z_{i_{k+1}}),k=0,\mathrm{},n1`$ | $`z_{k1}=\gamma _k(z_k),k=1,\mathrm{},n,z_n=z`$ |
| $`\varphi _{i_0,\mathrm{},i_n}(z_{i_n})dz_{i_n}^pd\overline{z}_{i_n}^q`$ | $`\varphi _{\gamma _1,\mathrm{},\gamma _n}(z)dz^pd\overline{z}^q`$ |
| (3.4.3) | (3.4.2) |
Similar provisions of course relate the deformed coordinates $`w_i`$ and elements of the deformed group $`\stackrel{~}{\mathrm{\Gamma }}`$. Any construction explicitly involving the map $`f`$ must take into account the equivariance property $`f\gamma =\stackrel{~}{\gamma }f`$ for any $`\gamma \mathrm{\Gamma }`$, where $`\stackrel{~}{\gamma }`$ is the corresponding element in the deformed group $`\stackrel{~}{\mathrm{\Gamma }}`$. We have relations entirely similar to (3.2.2) and (3.2.3) which can be found in ; for example
$$\frac{\stackrel{~}{\gamma }^{}f}{\gamma ^{}}f_z=f_z\gamma .$$
(3.4.4)
In order to handle the logarithm of (3.4.4) in the same way as we just did in the ฤech case (see (3.2.5)) we depart from . The problem is to relate $`\mathrm{log}(\gamma _1\gamma _2)^{}`$ and $`\mathrm{log}\gamma _1^{}\gamma _2+\mathrm{log}\gamma _2^{}`$ for any $`\gamma _1,\gamma _2\mathrm{\Gamma }`$, and similarly for $`\stackrel{~}{\mathrm{\Gamma }}`$. Instead of directly analyzing the branch-cuts (thus introducing an element of explicit dependence on the choice of the branches) we set
$`c_{\gamma _1,\gamma _2}`$ $`=\mathrm{log}\gamma _2^{}\mathrm{log}(\gamma _1\gamma _2)^{}+\mathrm{log}\gamma _1^{}\gamma _2,`$ (3.4.5a)
$`\stackrel{~}{c}_{\stackrel{~}{\gamma }_1,\stackrel{~}{\gamma }_2}`$ $`=\mathrm{log}\stackrel{~}{\gamma }_2^{}\mathrm{log}(\stackrel{~}{\gamma }_1\stackrel{~}{\gamma }_2)^{}+\mathrm{log}\stackrel{~}{\gamma }_1^{}\stackrel{~}{\gamma }_2,`$ (3.4.5b)
$`b_\gamma `$ $`=\mathrm{log}\stackrel{~}{\gamma }^{}f\mathrm{log}\gamma ^{}\mathrm{log}f_z\gamma +\mathrm{log}f_z.`$ (3.4.5c)
The numbers $`c_{\gamma _1,\gamma _2}`$, $`\stackrel{~}{c}_{\stackrel{~}{\gamma }_1,\stackrel{~}{\gamma }_2}`$ and $`b_\gamma `$ belong to $`(1)`$, and $`c`$, $`\stackrel{~}{c}`$ are cocycles with $`\stackrel{~}{c}=c+\delta b`$. Since $`\gamma ^{}`$ is the automorphy factor for $`TX`$, the geometric interpretation is that again $`c`$ represents $`c_1(TX)`$ . Alternatively, $`c`$ represents the Euler class of the $`S^1`$-bundle $`\mathrm{\Gamma }\backslash \mathrm{PSL}_2()X`$ (, see also ). Indeed, the first of equations (3.4.5) can be written in terms of rotation numbers:
$$c_{\gamma _1,\gamma _2}=2\left(w(\gamma _2)w(\gamma _1\gamma _2)+w(\gamma _1)\gamma _2\right),$$
where $`w\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)=\mathrm{arg}(cz+d)`$. More precisely, this is the Euler class of the $`^1`$-bundle obtained by letting $`\mathrm{PSL}_2()`$ act on the real projective line realized as the boundary of $``$ (see for details). Again, a similar discussion holds for $`\stackrel{~}{\mathrm{\Gamma }}`$ with the obvious changes.
As was shown in section 2.3, ฤech cohomology with respect to the cover $`X`$ is the same as group cohomology of $`\pi _1(X)\mathrm{\Gamma }`$ with values in the appropriate coefficients. Also it was noted there that $`X`$ is a good covering acyclic for fine sheaves, so that $`H^p(\pi _1(X),)H_{\mathrm{๐๐
}}^p(X)H^p(X,)`$. Similar arguments show that the double complex $`C^q(\mathrm{\Gamma },(3)_๐^p)`$ computes $`H_๐^{}(X,(3))`$.
The choice of the covering $`X`$ โ or, more generally, $`๐ปX`$ โ contains more information than simply using an abstract universal covering map $`UX`$: it includes the choice of a projective structure. Indeed, since the Schwarzian derivative of any Mรถbius transformation vanishes, any local section of the canonical projection would precisely be a system of projective charts for it.
It follows that when working with $`X`$ the explicit inclusion of projective connections becomes โ strictly speaking โ unnecessary. Indeed, these were not considered in . However, it is known that the effective action (that is, the class of the local Lagrangian cocycle) in higher genus is determined โ say, by the Universal Ward Identity โ only up to holomorphic quadratic differentials. Interpreting the latter as lifts of projective connections, the precise statement is that the effective action is determined up to the choice of a projective structure. In light of this observation, and also to keep a strict parallel with the ฤech formulation, we make this dependence on a generic projective connection explicit.<sup>8</sup><sup>8</sup>8Here the term projective connection is to be understood in the same way as in 3.3, i.e. as not necessarily holomorphic one. In this way we obtain a unified formalism consistent with the treatment of variations in section 4, where conditions on the projective connections will be enforced by the variational process.
Now we set out to write the correspondence:
| | $`๐ฐ_X`$ | $``$ |
| --- | --- | --- |
| $`(3,0)`$ | $`\omega _i(z_i)`$ | $`\omega (z)`$ |
| $`(2,1)`$ | $`\theta _{ij}(z_j)`$ | $`\theta _\gamma (z)`$ |
| $`(1,2)`$ | $`\mathrm{\Theta }_{ijk}(z_k)`$ | $`\mathrm{\Theta }_{\gamma _1,\gamma _2}(z)`$ |
| $`(0,3)`$ | $`m_{ijkl}`$ | $`m_{\gamma _1,\gamma _2,\gamma _3}`$ |
For the first two lines we start by translating (3.3.1) and (3.3.4), respectively:
$$\omega _\gamma (z)=\frac{^2f}{f}\mu dzd\overline{z}+2\mu hdzd\overline{z},$$
(3.4.6)
$$\theta _\gamma (z)=2\mu \frac{\gamma ^{\prime \prime }}{\gamma ^{}}d\overline{z}\left(\mathrm{log}(\stackrel{~}{\gamma }^{}f)+\mathrm{log}\gamma ^{}\right)d\mathrm{log}f+\mathrm{log}(\stackrel{~}{\gamma }^{}f)d\mathrm{log}\gamma ^{},$$
(3.4.7)
where $`h`$ is a smooth quadratic differential. In this way the last term of (3.4.6) is automorphic of weight $`(1,1)`$, hence it is killed by the coboundary operator. This would be consistent with a translation of (3.3.3). We stress (3.4.7) is a direct translation of the expression for the $`(2,1)`$ component *prior* to the computation of $`\delta \theta =d\mathrm{\Theta }`$. As before, the existence of $`\mathrm{\Theta }_{\gamma _1,\gamma _2}`$ is guaranteed by the vanishing of the analog of the symbol $`(TX,TX]`$ in holomorphic Deligne cohomology. This time, the tame symbol is represented by the cocycle
$$(\mathrm{log}\gamma _1^{}\gamma _2d\mathrm{log}\gamma _2^{},c_{\gamma _1,\gamma _2}\mathrm{log}\gamma _3^{},c_{\gamma _1,\gamma _2}c_{\gamma _3,\gamma _4})C^2(\mathrm{\Gamma },\underset{ยฏ}{\mathrm{\Omega }}^1())C^3(\mathrm{\Gamma },\underset{ยฏ}{๐ช}())C^4(\mathrm{\Gamma },(2)).$$
Since $`X`$ is a good cover, the quasi-isomorphism
$$(2)_{๐,\mathrm{โ๐๐}}^{}\stackrel{}{}\left(\underset{ยฏ}{๐ช}^{}()\stackrel{d\mathrm{log}}{}\underset{ยฏ}{\mathrm{\Omega }}^1()\right)[1]\stackrel{}{}/(2)^{}$$
is still in place by holomorphic Poincarรฉ lemma on $``$. Hence
$$^4(\mathrm{\Gamma },(2)_{๐,\mathrm{โ๐๐}}^{})H^3(\mathrm{\Gamma },^{})=0,$$
again, by obvious dimensional reasons. It follows that we can still introduce $`(\tau _\gamma )C^1(\mathrm{\Gamma },\underset{ยฏ}{\mathrm{\Omega }}^1())`$, $`(\varphi _{\gamma _1,\gamma _2})C^2(\mathrm{\Gamma },\underset{ยฏ}{๐ช}())`$ and $`(n_{\gamma _1,\gamma _2,\gamma _2})C^3(\mathrm{\Gamma },(2))`$ such that
$$(\mathrm{log}\gamma _1^{}\gamma _2d\mathrm{log}\gamma _2^{},c_{\gamma _1,\gamma _2}\mathrm{log}\gamma _3^{},c_{\gamma _1,\gamma _2}c_{\gamma _3,\gamma _4})=D(\tau _\gamma ,\varphi _{\gamma _1,\gamma _2},n_{\gamma _1,\gamma _2,\gamma _3}).$$
where various $`\gamma _i`$โs are used as place-holders for added clarity. Obviously, the treatment for the corresponding quantities depending on $`\stackrel{~}{\mathrm{\Gamma }}`$ is entirely similar. As a result, we can either compute the coboundary of (3.4.7) or simply translate (3.3.5) and repeat step by step what was done in section 3.3 to arrive at
$$\theta _\gamma =\theta _\gamma ^{\mathrm{๐๐๐}}\stackrel{~}{\tau }_\gamma +\tau _\gamma ,$$
(3.4.8)
with $`\theta _\gamma ^{\mathrm{๐๐๐}}`$ given by (3.4.7) and, finally:
$$\begin{array}{cc}\hfill \mathrm{\Theta }_{\gamma _1,\gamma _2}& =\stackrel{~}{\varphi }_{\gamma _1,\gamma _2}\varphi _{\gamma _1,\gamma _2}(\stackrel{~}{c}_{\gamma _1,\gamma _2}+c_{\gamma _1,\gamma _2})\mathrm{log}f\hfill \\ & \mathrm{log}\gamma _1^{}\gamma _2\mathrm{log}\stackrel{~}{\gamma }_{2}^{}{}_{}{}^{}+\stackrel{~}{c}_{\gamma _1,\gamma _2}\mathrm{log}(\gamma _1\gamma _2)^{}\hfill \end{array}$$
(3.4.9)
$$\begin{array}{cc}\hfill m_{\gamma _1,\gamma _2,\gamma _3}& =\stackrel{~}{n}_{\gamma _1,\gamma _2,\gamma _3}n_{\gamma _1,\gamma _2,\gamma _3}(\stackrel{~}{c}_{\gamma _1,\gamma _2}+c_{\gamma _1,\gamma _2})b_{\gamma _3}\hfill \\ & +c_{\gamma _1,\gamma _2\gamma _3}\stackrel{~}{c}_{\gamma _2,\gamma _3}c_{\gamma _1\gamma _2,\gamma _3}\stackrel{~}{c}_{\gamma _1,\gamma _2}.\hfill \end{array}$$
(3.4.10)
Therefore the analog of proposition 3.3.1 holds
###### Proposition 3.4.1.
The total cochain
$$\mathrm{\Omega }\stackrel{def}{=}2\pi \sqrt{1}(\omega ,\theta _\gamma ,\mathrm{\Theta }_{\gamma _1,\gamma _2},m_{\gamma _1,\gamma _2,\gamma _3}),$$
with $`\omega `$ given by the Polyakov form (3.4.6), represents a class in $`H_๐^3(X,(3))`$.
The action functional $`S[f]`$ is computed by evaluating $`\mathrm{\Omega }[f]`$ over the appropriate representative of $`[X]`$, which in this case would be a total cocycle in $`S_p()_\mathrm{\Gamma }B_q(\mathrm{\Gamma })`$ whose $`(2,0)`$ component can be taken as a fundamental domain $`F`$ for $`\mathrm{\Gamma }`$ in the form of a standard $`4g`$-gon, as detailed in .
## 4 Variation and projective structures
### 4.1 Variation
Here we compute the variation of the action functional $`S[f]`$ with respect to the dynamical field $`f`$, i.e. we compute its differential in field space. We denote by $`๐น`$ the variational operator โ the exterior differential in field space โ and we will use coordinates with respect to a good ฤech cover $`๐ฐ_X`$ whenever a local computation is required.
Since the dynamical field $`f`$ is a deformation map on $`X`$, we can either choose to allow variations that effectively deform the complex structure or restrict ourselves to the โtrivialโ ones โ deformations corresponding to vertical tangent vectors in the Earle-Eells fibration over the Teichmรผller space.
From (3.2.1) we get
$$f^{}(\kappa _{ij})=\frac{๐นf_i}{f_i}z_{ij}(z_{ij}^{})^1\frac{๐นf_j}{f_j},$$
(4.1.1)
where
$$\kappa =\{\kappa _{ij}\stackrel{def}{=}\frac{๐นw_{ij}}{w_{ij}^{}}\}$$
is the standard Kodaira-Spencer deformation cocycle, and $`f^{}(\kappa _{ij})=\kappa _{ij}f_j/f_j`$ is its pull-back. The condition $`[\kappa ]=0`$ in $`H^1(\stackrel{~}{X},\stackrel{~}{\mathrm{\Theta }})`$, where $`\stackrel{~}{\mathrm{\Theta }}`$ is the tangent sheaf of $`\stackrel{~}{X}`$, selects variations that leave the complex structure $`X`$ fixed. Specifically, if $`[\kappa ]=0`$ then it follows from (4.1.1) that $`๐นf_i/f_i`$ represents a smooth $`(1,0)`$-vector field on $`X`$ โ possibly after redefining it by a holomorphic coboundary for $`f^{}(\kappa _{ij})`$. Furthermore, the variation $`๐น\mu `$ of the corresponding Beltrami differential as a tangent vector to $`(X)`$ at $`\mu `$ is
$$๐น\mu =\overline{}_\mu \frac{๐นf}{f},$$
(4.1.2)
so the class $`[๐น\mu ]H_{\overline{}_\mu }^{(1,1)}(X)`$ corresponds to $`[\kappa ]`$ under the Dolbeault isomorphism.
In the sequel we shall confine ourselves to *vertical variations,* that is, to those with $`[\kappa ]=0`$. Then $`\frac{๐นf_i}{f_i}`$ defines a smooth vector field on $`X`$.
We start to compute the variation of the Lagrangian cocycle $`\mathrm{\Omega }`$ with respect to $`f`$. From a purely formal point of view, the calculation for the variation of the top form part proceeds as usual, where in each coordinate patch we have
$$๐น\omega _i=a_i+d\eta _i$$
with $`(a_i)\stackrel{ห}{C}^0(๐ฐ_X,\underset{ยฏ}{A}_X^2)`$ and $`(\eta _i)\stackrel{ห}{C}^0(๐ฐ_X,\underset{ยฏ}{A}_X^1)`$, where
$$a_i(f,๐นf)=2\overline{}_\mu \left(h_i\{f_i,z_i\}\right)\frac{๐นf_i}{f_i}dz_id\overline{z}_i.$$
Using the well-known identity
$$\overline{}_\mu \{f,z\}=^3\mu ,$$
where $`\mu =\mu (f)`$ and $`z`$ is a local coordinate on $`X`$ (the index $`i`$ is omitted here), we get
$`\overline{}_\mu (\{f,z\}h)`$ $`=(\overline{}\mu 2\mu )(\{f,z\}h)`$
$`=^3\mu (\overline{}\mu 2\mu )h`$
$`=^3\mu +2h\mu +h\mu \overline{}h`$
$`=๐_h\mu \overline{}h.`$
Here, for any smooth projective connection $`h๐ฌ(X)`$, $`๐_h`$ is the following third order differential operator:
$$๐_h=^3+2h+h.$$
(4.1.3)
It is well-known (see, e.g., ) that it has the property
$$๐_h:\underset{ยฏ}{A}_X^{1,l}\underset{ยฏ}{A}_X^{2,l},$$
for all $`l`$; in particular, $`๐_h`$ maps global forms of weight $`(1,l)`$ to global forms of weight $`(2,l)`$.
Thus the final expression for the variation of the top form term is,
$$a_i(f,๐นf)=2\left(\overline{}h_i๐_h\mu _i\right)\frac{๐นf_i}{f_i}dz_id\overline{z}_i.$$
(4.1.4)
Thanks to (3.3.2) and to the fact that $`๐_h`$ is a well defined map, $`a_i(f,๐นf)`$ is a well defined global $`2`$-form on $`X`$. The 1-form $`\eta _i`$ has the expression
$$\eta _i=๐น\mathrm{log}f_id\mathrm{log}f_i+2(\mathrm{log}f_i)dz_i\mathrm{}๐น\mu _i2(h_i\{f_i,z_i\})\frac{๐นf_i}{f_i}\left(dz_i+\mu _id\overline{z}_i\right),$$
(4.1.5)
where $`\mathrm{}`$ is the interior product between 1-forms and vectors.
The main point is that the term (4.1.4) alone constitutes the variation of the *whole* Lagrangian cocycle. Namely, we have
###### Theorem 4.1.1.
The variation of the total cocycle $`\mathrm{\Omega }[f]=2\pi \sqrt{1}(\omega _i,\theta _{ij},\mathrm{\Theta }_{ijk},m_{ijkl})`$ under vertical variation is given by the 2-form (4.1.4) up to a total coboundary in the Deligne complex. The variation of the action functional $`S[f]`$ is
$$๐นS[f]=2\pi \sqrt{1}_Xa(f,๐นf),$$
giving the following Euler-Lagrange equation
$$๐_h\mu \overline{}h=0.$$
(4.1.6)
We can give two different proofs of this theorem. One is more in keeping with the spirit of this work and uses the explicit form of $`\mathrm{\Omega }`$. The other is based only on Takensโ acyclicity theorem for the variational bicomplex and the formal machinery of descent equations. Although we present both, the second one will only be sketched here, as providing details for it would lead us to far afield.<sup>9</sup><sup>9</sup>9We plan to return to the topic from a more general point of view elsewhere.
###### First proof.
The procedure is to compute the variation of the various components of $`\mathrm{\Omega }`$ by applying $`๐น`$ to the descent equations. Start with $`\stackrel{ห}{\delta }๐น\omega _{ij}`$, that can be computed in two different ways: from equation $`\stackrel{ห}{\delta }\omega =d\theta `$, and from the variational relation $`๐น\omega =a+d\eta `$. Since $`a_i=a_j`$, we have
$$d(๐น\theta _{ij}\stackrel{ห}{\delta }\eta _{ij})=0,$$
and we deduce, using Poincarรฉ Lemma, that
$$๐น\theta _{ij}\stackrel{ห}{\delta }\eta _{ij}=d\lambda _{ij},$$
for $`(\lambda _{ij})\stackrel{ห}{C}^1(๐ฐ_X,\underset{ยฏ}{A}_X)`$. An explicit calculation using (3.3.7) and (4.1.5) confirms this relation with
$$\lambda _{ij}=2\frac{w_{ij}^{\prime \prime }}{w_{ij}^{}}f_j๐นf_j\left(\mathrm{log}w_{ij}^{}f_j+\mathrm{log}z_{ij}^{}\right)๐น\mathrm{log}f_j\stackrel{~}{\tau }_{ij}f_j๐นf_j.$$
(4.1.7)
The last term in this formula is obtained by varying the difference $`f^{}(\stackrel{~}{\tau }_{ij})\tau _{ij}`$, that enters equation (3.3.7). Clearly, the variation of $`\tau _{ij}`$ is zero and for the variation of $`f^{}(\stackrel{~}{\tau }_{ij})`$ we have
$`๐นf^{}(\stackrel{~}{\tau }_{ij})`$ $`=๐น\left(\stackrel{~}{\tau }_{ij}f_jdf_j\right)=๐น(\stackrel{~}{\tau }_{ij}f_j)df_j+\stackrel{~}{\tau }_{ij}f_j๐นdf_j`$
$`=\stackrel{~}{\tau }_{ij}^{}f_j๐นf_jdf_j+\stackrel{~}{\tau }_{ij}f_jd๐นf_j`$
$`=d\left(\stackrel{~}{\tau }_{ij}f_j๐นf_j\right),`$
since $`\stackrel{~}{\tau }_{ij}\mathrm{\Omega }^1(\stackrel{~}{U}_i\stackrel{~}{U}_j)`$ (see section 3.3.3).
Computing the coboundary of (4.1.7) yields
$$\begin{array}{c}\stackrel{ห}{\delta }\lambda _{ijk}=(\stackrel{~}{c}_{ijk}+c_{ijk})๐น\mathrm{log}f_k\hfill \\ \hfill (\mathrm{log}w_{ij}^{}f_j+\mathrm{log}z_{ij}^{})๐น\mathrm{log}w_{jk}^{}f_k(\stackrel{ห}{\delta }\stackrel{~}{\tau })_{ijk}f_k๐นf_k.\end{array}$$
On the other hand, the variation of (3.3.8) gives
$`๐น\mathrm{\Theta }_{ijk}`$ $`=\stackrel{~}{\varphi }_{ijk}^{}f_k๐นf_k(\stackrel{~}{c}_{ijk}+c_{ijk})๐น\mathrm{log}f_k\mathrm{log}z_{ij}^{}๐น\mathrm{log}w_{jk}^{}f_k`$
$`=\stackrel{ห}{\delta }\lambda _{ijk}+\mathrm{log}w_{ij}^{}f_j๐น(\mathrm{log}w_{jk}^{}f_k)+\stackrel{~}{\varphi }_{ijk}^{}f_k๐นf_k+(\stackrel{ห}{\delta }\stackrel{~}{\tau })_{ijk}f_k๐นf_k.`$
Using the first equation in (3.3.6):
$$d\stackrel{~}{\varphi }_{ijk}+(\stackrel{ห}{\delta }\tau )_{ijk}=\mathrm{log}w_{ij}^{}d\mathrm{log}w_{jk}^{},$$
we get
$$๐น\mathrm{\Theta }_{ijk}=\stackrel{ห}{\delta }\lambda _{ijk}.$$
Finally, putting it all together, we obtain
$$๐น\mathrm{\Omega }=(a_i)+D(\eta +\lambda ),$$
as wanted. โ
###### Second proof.
The $`2`$-form $`a_i`$ in the relation $`๐น\omega _i=a_i+d\eta _i`$ is a *source* form , hence it is uniquely determined by the de Rham class of $`\omega _i`$. Moreover, given a specific $`\omega _i`$, the form $`d\eta _i`$ is also determined (so $`\eta _i`$ is determined up to an exact form). Since $`\omega _j=\omega _i+d\theta _{ij}`$, we must have $`a_i=a_j`$ as both $`a_i`$ and $`a_j`$ are source forms for the *same* Lagrangian problem. Here the requirement that the variation be vertical is crucial in order to ensure that $`๐นf/f`$ glue as a geometric object โ a vector field on $`X`$. Therefore, from $`๐น\stackrel{ห}{\delta }\omega _{ij}=\stackrel{ห}{\delta }๐น\omega _{ij}`$, we get
$$๐น\theta _{ij}=\stackrel{ห}{\delta }\eta _{ij}+d\lambda _{ij}$$
by Poincarรฉ lemma. Proceeding in the same fashion we also get
$$d(๐น\mathrm{\Theta }_{ijk}\stackrel{ห}{\delta }\lambda _{ijk})=0.$$
Now, both $`๐น\mathrm{\Theta }_{ijk}`$ and $`\stackrel{ห}{\delta }\lambda _{ijk}`$ are forms of degree one in the field direction, i.e. they contain one variation. Takensโ acyclicity theorem asserts the variational bicomplex is acyclic in all degrees except the top one in the de Rham direction, provided the degree in the variational direction is at least one. Hence,
$$๐น\mathrm{\Theta }_{ijk}=\stackrel{ห}{\delta }\lambda _{ijk},$$
and we reach the same conclusion as in the previous proof. โ
### 4.2 Relative projective structures
Here we interpret of the Euler-Lagrange equation from the previous section through the principal $`๐ข(X)`$-bundle over the universal family of projective structures. First, we reformulate theorem 4.1.1 as follows
###### Theorem 4.2.1.
The Euler-Lagrange equation
$$\overline{}h_i=๐_h\mu _i$$
for the vertical variational problem is the condition that push-forward of the projective connection $`\{h_i\}`$ onto $`\stackrel{~}{X}`$ by the map $`f`$ is holomorphic.
###### Proof.
Indeed, the push-forward of $`h`$ is $`f_{}(h)=\{\stackrel{~}{h}f_i^1(f_i^1/w_i)^2\}`$, where $`\stackrel{~}{h}_i=h_i\{f_i,z_i\}`$. It is a projective connection on $`\stackrel{~}{X}`$ because of the transformation law
$$\stackrel{~}{h}_j\stackrel{~}{h}_iz_{ij}(z_{ij}^{})^2=\{w_i,w_j\}f_j(f_j)^2.$$
The Euler-Lagrange equation is equivalent to the equation $`\overline{}_\mu \stackrel{~}{h}_i=0`$, which is precisely the condition that projective connection $`f_{}(h)`$ is holomorphic on $`\stackrel{~}{X}`$. โ
It is well-known (see, e.g. ) that a holomorphic projective connection on $`X`$ determines a projective structure on $`X`$, and vice versa. The space of all projective structures on $`X`$ is an affine space modeled over $`H^0(X,\underset{ยฏ}{\mathrm{\Omega }}_X^2)`$ โ the vector space of holomorphic quadratic differentials on $`X`$.
For any holomorphic family $`CS`$ of Riemann surfaces parameterized by a complex manifold $`S`$, there is the holomorphic family $`P_S(C)S`$ of relative projective structures on $`C`$ . The fiber over $`sS`$ is the affine space of all (holomorphic) projective structures for $`C_s`$. We will be interested in the universal case $`S=๐ฏ(X)`$ and denote by $`๐ซ(X)`$ the universal family of relative projective structures. Following , consider the following pullback diagram
$$\begin{array}{ccc}๐ฎ(X)& & (X)\\ & & & & \\ ๐ซ(X)& & ๐ฏ(X)\end{array}$$
(4.2.1)
where the vertical arrows are principal $`๐ข(X)`$-bundles, and the horizontal ones are affine bundles with spaces affine over $`H^0(X_\mu ,\underset{ยฏ}{\mathrm{\Omega }}_{X_\mu }^2)`$ as fibers, $`\mu (X)`$. (The curve $`X_\mu `$ depends only on the class of $`\mu `$ modulo $`๐ข(X)`$ and so do its holomorphic objects.) Here $`๐ฎ(X)`$ is the space of all projective structures on $`X`$ holomorphic with respect to some complex structure determined by $`\mu (X)`$, without considering the quotient by $`๐ข(X)`$. Since every projective structure determines a complex structure, there is an obvious projection $`๐ฎ(X)(X)`$. As it follows from theorem 4.2.1,
$$๐ฎ(X)=\{(h,\mu )๐ฌ(X)\times (X)|๐_h\mu =\overline{}h\},$$
so that $`๐ฎ(X)`$ is the critical manifolds for the mapping $`A:๐ฌ(X)\times (X)^{}`$ (as well as for the map $`S:๐ฌ(X)\times (X)/(3)`$). The projection $`๐ฎ(X)(X)`$ is now just the projection on the second factor, and every fiber over $`\mu `$ in $`๐ฎ(X)`$ is indeed an affine space over the vector space $`H^0(X_\mu ,\underset{ยฏ}{\mathrm{\Omega }}_{X_\mu }^2)`$ (or rather its pull-back by $`f`$). Indeed, if $`(h,\mu )`$ and $`(h^{},\mu )`$ are two projective structures subordinated to $`\mu `$, then we have
$$๐_h\mu =\overline{}h\text{and}๐_h^{}\mu =\overline{}h^{},$$
which using the identity $`๐_h\mu \overline{}h=\overline{}_\mu (\{f,z\}h)`$ imply that
$$\overline{}_\mu (h^{}h)=0,$$
concluding that $`hh^{}`$ is a $`\mu `$-holomorphic quadratic differential.
The local meaning of the Euler-Lagrange equation โ the condition $`๐_h\mu =\overline{}h`$ โ is the following.
###### Lemma 4.2.2.
The operators $`\overline{}_\mu `$ and $`๐_h`$ commute if and only if the Euler-Lagrange equation is satisfied.
###### Proof.
Let $`v`$ be a local section of $`\underset{ยฏ}{A}_X^{1,0}`$. As a result of a direct calculation we have (omitting the coordinate index $`i`$)
$$๐_h\overline{}_\mu v\overline{}_\mu ๐_hv=L_v(๐_h\mu \overline{}h),$$
(4.2.2)
where $`L_v=v+2v`$ is the Lie derivative operator on $`\underset{ยฏ}{A}_X^{2,1}`$. Thus the โifโ part is clear. For the โonly ifโ part, assume the RHS of (4.2.2) is zero for all $`v`$. Therefore, if we consider $`fv`$ for any local $`f`$, then we must have $`v(f)(๐_h\mu \overline{}h)=0`$, implying (4.1.6). โ
We conclude that $`๐ฎ(X)`$ is the geometric locus where the commutativity condition $`๐_h\overline{}_\mu =\overline{}_\mu ๐_h`$ is satisfied. Then we can consider $`๐_h`$ as a map between two augmented Dolbeault complexes (where $`\underset{ยฏ}{\mathrm{\Theta }}_\mu `$ and $`\underset{ยฏ}{\mathrm{\Omega }}_{X_\mu }^2`$ are actually pull-backs of the corresponding sheaves from $`X_\mu `$ to $`X`$ by the map $`f(\mu )`$):
$$\begin{array}{ccccccccc}0& & \underset{ยฏ}{\mathrm{\Theta }}_\mu & \stackrel{ฤฑ}{}& \underset{ยฏ}{A}_X^{1,0}& \stackrel{\overline{}_\mu }{}& \underset{ยฏ}{A}_X^{1,1}& & 0\\ & & ๐_h& & ๐_h& & ๐_h& & \\ 0& & \underset{ยฏ}{\mathrm{\Omega }}_{X_\mu }^2& \underset{ฤฑ}{}& \underset{ยฏ}{A}_X^{2,0}& \underset{\overline{}_\mu }{}& \underset{ยฏ}{A}_X^{2,1}& & 0\end{array}$$
(4.2.3)
where the morphism $`\underset{ยฏ}{\mathrm{\Theta }}_\mu \stackrel{๐_h}{}\underset{ยฏ}{\mathrm{\Omega }}_{X_\mu }^2`$ is now the usual third order $`\mu `$-holomorphic operator , also familiar from the theory of the KdV equation . It fits into the exact sequence
$$\begin{array}{ccccccccc}0& & \underset{ยฏ}{V}_X(h)& \stackrel{ฤฑ}{}& \underset{ยฏ}{\mathrm{\Theta }}_\mu & \stackrel{๐_h}{}& \underset{ยฏ}{\mathrm{\Omega }}_{X_\mu }^2& & 0\end{array}$$
(4.2.4)
where $`\underset{ยฏ}{V}_X(h)`$ is a rank three local system depending on the projective structure $`h`$ โ a locally constant sheaf on $`X`$. Actually, it is the sheaf of polynomial vector fields of degree not greater than two in the coordinates adapted to $`(h,\mu )`$. Passing to cohomology, we get:
$$0H^0(X_\mu ,\underset{ยฏ}{\mathrm{\Omega }}_{X_\mu }^2)H^1(X,\underset{ยฏ}{V}_X(h))H^1(X_\mu ,\mathrm{\Theta }_\mu )0$$
(4.2.5)
According to the theorem of Hubbard , sequence (4.2.5) is isomorphic to the tangent bundle sequence for the relative projective structure $`๐ซ(X)๐ฏ(X)`$ at $`(h,\mu )`$. Furthermore, the usual machinery of local systems shows that $`H^1(X,\underset{ยฏ}{V}_X(h))`$ is isomorphic to the Eichler cohomology group $`H^1(\pi _1(X,p),๐ฝ(h)_p)`$, where $`๐ฝ(h)_p`$ is the stalk of $`\underset{ยฏ}{V}_X(h)`$ over the point $`p`$. The proof that this coincides with the classical Eichler cohomology (see ), can be obtained by lifting everything to the universal cover $``$ of $`X`$ and using factors of automorphy (see for further details).
On the other hand, from our description of $`๐ฎ(X)`$ we have
$$T_{(h,\mu )}๐ฎ(X)=\{(\dot{h},\dot{\mu })A^{2,0}(X)\times A^{1,1}(X)|๐_h\dot{\mu }=\overline{}_\mu \dot{h}\}$$
(4.2.6)
and the RHS can be written as the fiber product
$$A^{2,0}(X)\times _{A^{2,1}(X)}A^{1,1}(X)$$
with respect to the pair of maps $`\overline{}_\mu `$ and $`๐_h`$. For vertical โ along the fiber of $`๐ฎ(X)(X)`$ โ tangent vectors to $`๐ฎ(X)`$ at $`(h,\mu )`$ we have
$$(\dot{h},\dot{\mu })=(๐_hv,\overline{}_\mu v),$$
where $`vA^{1,0}(X)`$ is the infinitesimal generator. This pair clearly satisfies the condition in (4.2.6), since $`๐ฎ(X)`$ is the geometric locus of the commutativity condition. Thus the map sending $`v(๐_hv,\overline{}_\mu v)`$ describes the vertical tangent bundle of $`๐ฎ(X)(X)`$. Therefore, if $`[h,\mu ]`$ denotes the class of $`(h,\mu )`$, we have for the vertical tangent space to $`๐ฎ(X)`$ at $`[h,\mu ]`$:
$$T_{V,[h,\mu ]}๐ฎ(X)\left(A^{2,0}(X)\times _{A^{2,1}(X)}A^{1,1}(X)\right)/(๐_h,\overline{}_\mu )(A^{1,0}(X)),$$
(4.2.7)
which obviously projects onto $`H_{\overline{}_\mu }^{1,1}(X)H^1(X_\mu ,\mathrm{\Theta }_\mu )`$. Now, this is just the $`C^{\mathrm{}}`$ image of the Eichler cohomology description of the tangent sheaf to the relative projective structure $`๐ซ(X)๐ฏ(X)`$ and we have the following
###### Proposition 4.2.3.
The differential geometric description of the tangent space to $`๐ฏ(X)`$ at the class of $`(h,\mu )`$ as given by (4.2.7) coincides with the algebraic description given by the Eichler cohomology group $`H^1(X,\underset{ยฏ}{V}_X(h))`$.
###### Proof.
Consider the cone of $`๐_h:\underset{ยฏ}{A}_X^{1,}\underset{ยฏ}{A}_X^{2,}`$:
$$\underset{ยฏ}{C}_X^{}:0\underset{ยฏ}{A}_X^{1,0}\stackrel{\overline{}_\mu ๐_h}{}\underset{ยฏ}{A}_X^{1,1}\underset{ยฏ}{A}_X^{2,0}\stackrel{๐_h\overline{}_\mu }{}\underset{ยฏ}{A}_X^{2,1}0$$
Its cohomology sheaf complex equals $`\underset{ยฏ}{V}_X(h)`$, thus by standard homological algebra arguments (see, e.g. ) one has $`^1(X,\underset{ยฏ}{C}_X^{})=H^1(X,\underset{ยฏ}{V}_X(h))`$ and from the canonical sequence
$$0\underset{ยฏ}{A}_X^{2,}[1]\underset{ยฏ}{C}_X^{}\underset{ยฏ}{A}_X^{1,}0$$
one gets (4.2.5). On the other hand, the RHS of (4.2.7) is the first cohomology group of the complex
$$0A^{1,0}(X)\stackrel{\overline{}_\mu ๐_h}{}A^{1,1}(X)A^{2,0}(X)\stackrel{๐_h\overline{}_\mu }{}A^{2,1}(X)0$$
which is equal to the first term
$${}_{}{}^{\mathrm{}}E_{1}^{p,q}\stackrel{ห}{H}^q(X,\underset{ยฏ}{C}_X^p)=\{\begin{array}{cc}C^p(X)\hfill & q=0,\hfill \\ 0\hfill & q>0.\hfill \end{array}$$
of the spectral sequence computing $`^{}(X,\underset{ยฏ}{C}_X^{})`$. โ
### 4.3 Geometry of the vertical variation
Here we consider functional $`A[f]`$ as as map $`A:๐ฌ(X)\times (X)^{}`$, where $`๐ฌ(X)`$ is the affine space of all $`C^{\mathrm{}}`$ projective connections on $`X`$ and $`(X)`$ is the total space of the Earle-Eells fibration.
By theorem 4.2.1, the critical manifold for $`A[f]`$ coincides with $`๐ฎ(X)`$. Considering critical values of $`A`$ (โon shellโ condition) leads to the function $`A:๐ฎ(X)^{}`$, where $`A(h,\mu )=\mathrm{\Omega }[h,\mu ],\mathrm{\Sigma }_m`$. Since $`๐ฎ(X)`$ is a principal $`๐ข(X)`$-bundle over $`๐ซ(X)`$, it is interesting to analyze the behavior of $`A`$ under the $`๐ข(X)`$-action. It is given by the following
###### Lemma 4.3.1.
The directional derivative of the action functional $`A`$ for the vertical tangent vector $`(๐_hv,\overline{}_\mu v)`$ to $`๐ฎ(X)`$ at $`(h,\mu )`$, where $`vA^{1,0}(X)`$, is given by
$$4\pi \sqrt{1}_X\mu ๐_hvA.$$
(4.3.1)
###### Proof.
We just repeat the computation of the vertical variation with additional term $`2\mu ๐นhdzd\overline{z}`$, where $`๐นh=๐_hv`$. Since the main term, given by $`2(\overline{}h๐_h\mu )vdzd\overline{z}`$ vanishes โon shellโ, this proves the result. โ
Formula (4.3.1) defines a function $`c:๐ฎ(X)\left(\mathrm{Lie}๐ข(X)\right)^{}`$ by assigning to the pair $`(h,\mu )`$ a linear functional on $`\mathrm{Lie}๐ข(X)A^{1,0}(X)`$ as follows:
$$v2_X\mu ๐_hv.$$
(4.3.2)
Equivalently, $`c`$ is a 1-cochain over $`\mathrm{Lie}๐ข(X)`$ with values in functions over $`๐ฎ(X)`$ with left $`\mathrm{Lie}๐ข(X)`$-action.
###### Proposition 4.3.2.
The 1-cochain $`c`$ is a 1-cocycle.
###### Proof.
For $`v,wA^{1,0}(X)\mathrm{Lie}๐ข(X)`$ we have
$$\delta c(v,w)=vc(w)wc(v)c([v,w])$$
where $`c(u):๐ฎ(X)`$ is the function
$$c(u)(h,\mu )=2_X\mu ๐_hu.$$
Using the infinitesimal action,
$$vc(w)(h,\mu )=2_X(\overline{}_\mu v๐_hw+\mu L_v(๐_hw),$$
we get
$$\begin{array}{cc}\hfill (\delta c)(h,\mu )=2_X(\overline{}_\mu v๐_hw& +\mu L_v(๐_hw\hfill \\ \hfill \overline{}_\mu w๐_hv& \mu L_w(๐_hv)\mu ๐_hL_vw).\hfill \end{array}$$
where $`L_v=v+2v`$ is the Lie derivative on $`A^2(X)`$, and the Lie bracket in $`A^{1,0}(X)`$ is the usual vector field Lie bracket: $`[v,w]=L_vw=(vwwv)`$. Using the identity
$$L_v(๐_hw)L_w(๐_hv)๐_hL_v(w)=0,$$
we are left with
$`(\delta c)(v,w)(h,\mu )`$ $`=2{\displaystyle _X}\left(\overline{}_\mu v๐_hw\overline{}_\mu w๐_hv\right)`$
$`=2{\displaystyle _X}v\left(๐_h\overline{}_\mu w\overline{}_\mu ๐_hw\right)`$
$`=0,`$
because of the commutativity condition and the skew-symmetry of the operator $`๐_h`$. โ
## Appendix A Appendix
### A.1 Cones
Recall that for a map $`u:๐ ^{}๐ก^{}`$ the *cone* $`๐ข_u^{}`$ of $`u`$ is the complex:
$$๐ข_u^{}=๐ ^{}[1]๐ก^{}$$
with differential
$$d(a,b)=(da,u(a)+db).$$
The cone fits into the exact sequence:
$$0๐ก^{}๐ข_u^{}๐ ^{}[1]0.$$
If the map $`u`$ is injective, this is the same as the cokernel of $`u`$ (up to a shift in the resulting exact cohomology sequence).
For the Deligne complex, we often find that the equivalent definition of $`(p)_๐^{}`$ is
$$(p)_๐^{}=\mathrm{Cone}\left((p)F^p(\underset{ยฏ}{A})_M^{}\stackrel{ฤฑศท}{}\underset{ยฏ}{A}_M^{}\right)[1],$$
(A.1.1)
where $`ศท:F^p(\underset{ยฏ}{A})_M^{}\underset{ยฏ}{A}_M^{}`$ is the Hodge-Deligne filtration (*filtration bรชte*), that is, the $`n`$-th sheaf of $`F^p(\underset{ยฏ}{A})_M^{}`$ is $`\underset{ยฏ}{A}_M^n`$ if $`np`$, and zero otherwise.
Briefly, the equivalence is shown as follows. The cone in (A.1.1) is equal to
$$\mathrm{Cone}\left((p)\mathrm{Cone}(F^p(\underset{ยฏ}{A})_M^{}\underset{ยฏ}{A}_M^{})\right)[1].$$
The inner cone can clearly be replaced by the cokernel of the inclusion map, namely the (sharp) truncation $`\tau ^{p1}\underset{ยฏ}{A}_M^{}`$ of the de Rham complex. Thus we have
$$\mathrm{Cone}\left((p)\tau ^{p1}\underset{ยฏ}{A}_M^{}\right)[1],$$
which equals $`(p)_๐^{}`$ as defined in the main text.
### A.2 Fundamental class
We want to collect here some technical facts and computations related to the construction of a representative of the fundamental class $`[M]`$, that are not strictly necessary in the main body of this paper.
Recall that we work with the double complex
$$๐ฒ_{p,q}=S_p(N_q(UX)),$$
where $`N_{}(UX)`$ is the nerve of the covering $`UX`$, and $`S_{}`$ is the singular simplices functor.
#### A.2.1
We saw in the main text, sec. 2.4, that when $`๐ฒ_{p,}`$ *resolves* $`S_p(M)`$ for any fixed $`p`$, the total homology of $`๐ฒ_,`$ is equal to $`H_{}(M,)`$. By definition, this condition is that $`H_0(S_p(N_{}(U)))S_p(M)`$ and $`H_q(S_p(N_{}(U)))=0`$ for $`q>0`$. Then the isomorphism $`H_{}(M,)H_{}(\mathrm{Tot}๐ฒ)`$ can be easily obtained by carefully lifting a cocycle in $`S_{}(M)`$ to a total cocycle in $`๐ฒ_,`$.<sup>10</sup><sup>10</sup>10See, e.g., . Details for this calculation can be found in the appendix of . More concisely, we have $`{}_{}{}^{\mathrm{}}E_{p,q}^{1}=H_q(S_p(N_{}(U)))=0`$ for $`q>0`$ (the spectral sequence collapses) and at the next step one has $`{}_{}{}^{\mathrm{}}E_{p,0}^{2}={}_{}{}^{\mathrm{}}E_{p,0}^{\mathrm{}}H_p(S_{}(M))=H_p(M,)`$, as wanted.
These requirements are met for a ฤech covering $`๐ฐ_M`$, where a contracting homotopy for $`S_p(N_{}(๐ฐ_M))`$ can be constructed explicitly (see also , appendix on the de Rham theorem). Indeed, one can easily show that $`H_0(S_p(N_{}(๐ฐ_M))S_p(M)`$ by applying $`S_p()`$ to the sequence $`\mathrm{}N_1(๐ฐ_M)N_0(๐ฐ_M)M`$. The resulting maps are $`_i\sigma _i_i\sigma _i`$ and $`_{ij}\sigma _{ij}_i(_j(\sigma _{ji}\sigma _{ij}))`$, so the composition is zero. Moreover, if $`_i\sigma _i=0`$, for any pair of indices $`ij`$, we must have $`\sigma _i|_{U_{ij}}+\sigma _j|_{U_{ij}}=0`$, so that $`\sigma _i|_{U_{ij}}\sigma _j|_{U_{ij}}=\sigma _i|_{U_{ij}}\sigma _i|_{U_{ij}}=^{\prime \prime }\sigma _i|_{U_{ij}}`$, proving the claim.
Similarly, if $`UM`$ is a regular covering with $`G=\mathrm{Deck}(U/M)`$ acting on the right on $`U`$, then $`๐ฒ_{p,0}=S_p(N_0(U))S_p(U)`$ is a free (right) $`G`$-module , so that $`S_p(N_{}(U))S_p(U)_GB_{}(G)`$ resolves $`S_p(U)_GS_p(M)`$ hence
$$H_q(S_p(N_{}(UM)))\{\begin{array}{cc}S_p(M)\hfill & q=0\hfill \\ 0\hfill & q>0,\hfill \end{array}$$
as wanted.
#### A.2.2
Since $`๐ฒ_,`$ is a double complex, it is well known that its associated total complex can be filtered in two ways โ with respect to either $`p`$ or $`q`$. Filtering over the second index of $`๐ฒ_{p,q}=S_p(N_q(U))`$ yields the second spectral sequence with
$${}_{}{}^{\mathrm{}\mathrm{}}E_{p,q}^{1}H_p^{^{}}(๐ฒ_{,q})H_p(S_{}(N_q(U)).$$
Although not required in the following it is interesting to see when and whether this latter sequence also degenerates, like the other one. In other words, we want to consider the case when for fixed $`q`$ the complex $`๐ฒ_{,q}`$ is acyclic in degree $`>0`$.
###### Assumption.
The covering $`UM`$ is *good*, that is, each $`N_q(U)=U\times _M\mathrm{}\times _MU`$ is contractible, hence is acyclic for the singular simplices functor.
###### Remark A.2.1.
The assumption on $`UM`$ guarantees the de Rham complex is a resolution of $``$, so the second cohomological spectral sequence $`H^p(\stackrel{ห}{C}^q(U;\underset{ยฏ}{A}^{}))`$ degenerates and the total cohomology equals $`\stackrel{ห}{H}^q(U;)`$.
By virtue of the assumption, $`{}_{}{}^{\mathrm{}\mathrm{}}E_{}^{1}`$ is computed as
$${}_{}{}^{\mathrm{}\mathrm{}}E_{q,p}^{1}\{\begin{array}{cc}<N_q๐_U>\hfill & p=0\hfill \\ 0\hfill & p>0,\hfill \end{array}$$
where $`N_q๐_U`$ is the set of connected components of $`N_q(U)`$ and $`<N_q๐_U>`$ is the abelian group generated by $`N_q๐_U`$. This follows from the fact that $`H_0`$ gives us a factor $``$ for every connected component of $`N_q(U)`$. These connected components arrange into a simplicial set $`N_{}๐_U`$, where the face maps are induced by the face maps of the nerve $`N_{}(U)`$, specifying where every component goes. Thus $`N_{}๐_U`$ expresses the pure combinatorics of the covering. Since the spectral sequence collapses, the total homology is equal to
$${}_{}{}^{\mathrm{}\mathrm{}}E_{q,0}^{2}={}_{}{}^{\mathrm{}\mathrm{}}E_{q,0}^{\mathrm{}}H_q(<N_{}R_U>)$$
and (see )
$$H_q(<N_{}R_U>)H_q(|N_{}R_U|),$$
where $`||`$ is the geometric realization of $`N_q๐_U`$, namely, the CW-complex obtained by putting in a standard $`q`$-simplex $`\mathrm{\Delta }^q`$ for each element in $`N_q๐_U`$ and gluing them together according to the face maps. Therefore, for a good covering the three homologies are equal:
$$H_q(\mathrm{Tot}๐ฒ_,)H_q(M,)H_q(|N_{}๐_U|).$$
In our concrete examples, an ordinary ฤech covering is good if all $`U_{i_0}\mathrm{}U_{i_q}`$ are contractible. In this case, to compute $`H_0^{^{}}(S_{}(N_q๐ฐ_M))`$ we must assign a $``$ factor to each $`U_{i_0}\mathrm{}U_{i_q}`$. Following , denote $`U_{i_0}\mathrm{}U_{i_q}`$ as a generator in this group by the symbol $`\mathrm{\Delta }_{i_0,\mathrm{},i_q}`$, so that $`<N_q๐_U>=_{i_0,\mathrm{},i_q}\mathrm{\Delta }_{i_0,\mathrm{},i_q}`$ and $`N_q๐_U=\{\mathrm{\Delta }_{i_0,\mathrm{},i_q}\}`$. Therefore, $`N_{}๐_U`$ represents the abstract nerve of the open cover and $`|N_{}๐_U|`$ is the CW-complex obtained by replacing each $`\mathrm{\Delta }_{i_0,\mathrm{},i_q}`$ โ in other words, each non void intersection โ by a standard $`q`$-simplex and gluing them according to the face maps of $`N_{}๐_U`$.
On the other hand, if $`UM`$ is a $`G`$-covering, then according to 2.3 $`N_q(U)=U\times G^q`$, and it is good if $`U`$ is contractible. Thus $`<N_{}๐_U>_GB_{}(G)`$, so that
$$H_q(<N_{}๐_U>)H_q(G;)H_q(BG,),$$
where $`BG=|N_{}๐_U|`$ is the classifying space of $`G`$, where in this case $`N_q๐_U=G^q`$ for $`q1`$ and $`N_0๐_U=\mathrm{point}`$.
#### A.2.3
Let us return to the main problem of representing the fundamental class of $`X`$ as a total cycle in the double complex $`๐ฒ_{p,q}`$. If the sequence $`0S_p(X)S_p(N_{}(U))`$ is exact, then there exists a splitting $`\tau :S_p(X)S_p(N_{}(U))`$, i.e. the map $`\tau `$ satisfies $`ฯต\tau =\mathrm{id}_{S_p(X)}`$. In other words, $`\tau `$ is the first step of an explicit contracting homotopy for $`S_p(N_{}(U))`$. Then a cycle representing $`[X]`$ can be produced by lifting $`X`$ via $`\tau `$ and completing $`\tau (X)`$ to a total cycle using the standard descent argument.
In the concrete examples we have been looking at, this can be done as follows. The case where $`U`$ is a regular $`G`$-covering can be handled by starting from a fundamental domain $`F`$ for the action of $`G`$ on $`U`$, where we regard $`F`$ as an element of degree $`(p,0)`$ in $`๐ฒ_{p,0}S_p(U)_GB_0(G)S_p(U)`$. Full details are spelled out in . If $`U`$ comes from an ordinary ฤech covering $`๐ฐ_X`$, we first replace $`S_p(X)`$ by $`๐ฐ_X`$-small simplices:
$$0S_p^๐ฐ(X)S_p(N_{}๐ฐ_X),$$
where the $`๐ฐ_X`$-small simplices are those whose support is contained in the open cover $`๐ฐ_X=\{U_i\}`$. Second, write $`X=_i\sigma _i`$, where all $`\sigma _i`$ are $`๐ฐ_X`$-small, and set $`\tau (X)=_i\sigma _i\mathrm{\Delta }_i\stackrel{def}{=}\mathrm{\Sigma }_0๐ฒ_{p,0}`$. Since
$$ฯต(^{}\mathrm{\Sigma }_0)=^{}ฯต\mathrm{\Sigma }_0=^{}ฯต\tau (X)=^{}X0,$$
by the standard argument there exist $`\mathrm{\Sigma }_1,\mathrm{\Sigma }_2,\mathrm{},\mathrm{\Sigma }_p`$, with $`\mathrm{\Sigma }_k๐ฒ_{pk,k},k=1,\mathrm{},p,`$ such that
$$^{}\mathrm{\Sigma }_0=^{\prime \prime }\mathrm{\Sigma }_1,\mathrm{},^{}\mathrm{\Sigma }_{q1}=^{\prime \prime }\mathrm{\Sigma }_q,\mathrm{},^{}\mathrm{\Sigma }_p=0.$$
This schema can be implemented in a fairly explicit way using a map $`h:N_{}๐_๐ฐS_{}^๐ฐ(X)`$ constructed in (Th. 13.4, proof) to realize the nerve of a covering. Of course, our case of interest here is $`p=2`$.
In order to describe $`h`$ we shall need the barycentric decomposition $`N_{}\stackrel{~}{๐}_๐ฐ`$ of $`N_{}๐_๐ฐ`$ (see for a more complete explanation). For any finite subset $`\tau `$ of the index set $`I`$ denote $`U_\tau =_{i\tau }U_i`$, and let:
$$N_0\stackrel{~}{๐}_๐ฐ=\underset{\tau I}{}\{U_\tau \}$$
$$N_1\stackrel{~}{๐}_๐ฐ=\underset{\tau _0\tau _1I}{}\{U_{\tau _1}U_{\tau _0}\}$$
$$\mathrm{}$$
$$N_q\stackrel{~}{๐}_๐ฐ=\underset{\tau _0\mathrm{}\tau _qI}{}\{U_{\tau _q}\mathrm{}U_{\tau _0}\}.$$
In order to construct the mapping $`h`$, assign to each $`U_\tau `$ a point $`v_\tau U_\tau `$, to any inclusion $`U_{\tau _1}U_{\tau _0}`$ a path from $`v_{\tau _0}`$ to $`v_{\tau _1}`$, and to $`U_{\tau _2}U_{\tau _1}U_{\tau _0}`$ the cone from $`v_{\tau _0}`$ to the path from $`v_{\tau _1}`$ to $`v_{\tau _2}`$, which is of course a 2-simplex. Denote by $`\mathrm{\Delta }(v_{\tau _0})`$, $`\mathrm{\Delta }(v_{\tau _0},v_{\tau _1})`$ and $`\mathrm{\Delta }(v_{\tau _0},v_{\tau _1},v_{\tau _2})`$ the $`0`$, $`1`$ and $`2`$-simplices so obtained. Observe how the simplices constructed in this way inherit an orientation from the natural one on the barycentric decomposition $`N_{}\stackrel{~}{๐}_๐ฐ`$; this is the main reason for using $`N_{}\stackrel{~}{๐}_๐ฐ`$ in place of $`N_{}๐_๐ฐ`$. So, for example, $`\mathrm{\Delta }(v_{\tau _0},v_{\tau _1},v_{\tau _2})`$ has the orientation induced by the order $`v_{\tau _0}v_{\tau _1}v_{\tau _2}`$ associated to the inclusion $`\tau _0\tau _1\tau _2`$. The typical situation for the indices $`i,j,k`$ looks as in figure 1: to the index sets $`i`$, $`ij`$ and $`ijk`$ correspond the points $`v_i`$, $`v_{ij}`$ and $`v_{ijk}`$ in $`U_i`$, $`U_{ij}`$ and $`U_{ijk}`$, respectively. Then $`\mathrm{\Delta }(v_i,v_{ij})`$ is the $`1`$-simplex joining $`v_i`$ and $`v_{ij}`$, $`\mathrm{\Delta }(v_{ij},v_{ijk})`$ the one joining $`v_{ij}`$ and $`v_{ijk}`$, and so on.
After these preparations, define an element $`\mathrm{\Sigma }_0`$ in $`๐ฒ_{2,0}`$ as
$$\mathrm{\Sigma }_0=\underset{iI}{}\mathrm{st}(v_i)\mathrm{\Delta }_i,$$
where
$$\mathrm{st}(v_i)=\underset{j,k:\mathrm{\Delta }_{ijk}0}{}ฯต_{ijk}\left(\mathrm{\Delta }(v_i,v_{ij},v_{ijk})\mathrm{\Delta }(v_i,v_{ik},v_{ijk})\right)$$
is the star of the vertex $`v_i`$, and $`ฯต_{ijk}=\pm 1`$ according to whether the order of the triple $`i,j,k`$ agrees the orientation or not, namely whether the order $`ijk`$ is the same as the cyclic (counterclockwise) order around the vertex $`v_{ijk}`$. Recall that $`\mathrm{\Delta }_\tau `$ is the symbol corresponding to $`U_\tau `$, when considered as a generator in the abelian group generated by the nerve, as in A.2.2. Rewriting $`\mathrm{\Sigma }_0`$ as
$`{\displaystyle \underset{iI}{}}\mathrm{st}(v_i)\mathrm{\Delta }_i={\displaystyle \underset{i,j,k}{}}ฯต_{ijk}\{`$ $`\left(\mathrm{\Delta }(v_i,v_{ij},v_{ijk})\mathrm{\Delta }(v_i,v_{ik},v_{ijk})\right)\mathrm{\Delta }_i`$
$``$ $`\left(\mathrm{\Delta }(v_j,v_{ij},v_{ijk})\mathrm{\Delta }(v_j,v_{jk},v_{ijk})\right)\mathrm{\Delta }_j`$
$`+`$ $`(\mathrm{\Delta }(v_k,v_{ik},v_{ijk})\mathrm{\Delta }(v_k,v_{jk},v_{ijk}))\mathrm{\Delta }_k\},`$
where $`_{i,j,k}`$ means sum over triples of indices in $`I`$, its first differential is:
$`^{}\mathrm{\Sigma }_0`$ $`={\displaystyle \underset{i,j,k}{}}ฯต_{ijk}\{\mathrm{\Delta }(v_{ik},v_{ijk})(\mathrm{\Delta }_k\mathrm{\Delta }_i)`$
$`\mathrm{\Delta }(v_{ij},v_{ijk})(\mathrm{\Delta }_j\mathrm{\Delta }_i)\mathrm{\Delta }(v_{jk},v_{ijk})(\mathrm{\Delta }_k\mathrm{\Delta }_j)\}`$
$`+{\displaystyle \underset{i,j,k}{}}ฯต_{ijk}\{\mathrm{\Delta }(v_i,v_{ij})\mathrm{\Delta }_i\mathrm{\Delta }(v_i,v_{ik})\mathrm{\Delta }_i\mathrm{\Delta }(v_j,v_{ij})\mathrm{\Delta }_j`$
$`+\mathrm{\Delta }(v_j,v_{jk})\mathrm{\Delta }_j+\mathrm{\Delta }(v_k,v_{ik})\mathrm{\Delta }_k\mathrm{\Delta }(v_k,v_{jk})\mathrm{\Delta }_k\}.`$
The last sum is easily seen to be zero, while the first can be rewritten as $`^{\prime \prime }\mathrm{\Sigma }_1`$ for the following element in $`๐ฒ_{1,1}`$:
$$\mathrm{\Sigma }_1=\underset{i,j,k}{}ฯต_{ijk}\left\{\mathrm{\Delta }(v_{ik},v_{ijk})\mathrm{\Delta }_{ik}\mathrm{\Delta }(v_{ij},v_{ijk})\mathrm{\Delta }_{ij}\mathrm{\Delta }(v_{jk},v_{ijk})\mathrm{\Delta }_{jk}\right\}$$
Again, computing the first differential gives
$`^{}\mathrm{\Sigma }_1`$ $`={\displaystyle \underset{i,j,k}{}}ฯต_{ijk}\left\{v_{ijk}(\mathrm{\Delta }_{ik}\mathrm{\Delta }_{ij}\mathrm{\Delta }_{jk})\right\}`$
$`+{\displaystyle \underset{i,j,k}{}}ฯต_{ijk}\left\{v_{ij}\mathrm{\Delta }_{ij}v_{ik}\mathrm{\Delta }_{ik}+v_{jk}\mathrm{\Delta }_{jk}\right\},`$
with the last sum being identically zero. The first term can be rewritten as $`^{\prime \prime }\mathrm{\Sigma }_2`$, where
$$\mathrm{\Sigma }_2=\underset{i,j,k}{}ฯต_{ijk}v_{ijk}\mathrm{\Delta }_{ijk}.$$
Finally, the total chain $`\mathrm{\Sigma }\mathrm{\Sigma }_0+\mathrm{\Sigma }_1\mathrm{\Sigma }_2`$ is a cycle, $`\mathrm{\Sigma }=0`$, and we have the following expression for the representative of the fundamental class of $`X`$ in the double complex:
$`\mathrm{\Sigma }`$ $`={\displaystyle \underset{i,j,k}{}}ฯต_{ijk}\{(\mathrm{\Delta }(v_i,v_{ij},v_{ijk})\mathrm{\Delta }(v_i,v_{ik},v_{ijk}))\mathrm{\Delta }_i`$
$`\left(\mathrm{\Delta }(v_j,v_{ij},v_{ijk})\mathrm{\Delta }(v_j,v_{jk},v_{ijk})\right)\mathrm{\Delta }_j`$
$`+(\mathrm{\Delta }(v_k,v_{ik},v_{ijk})\mathrm{\Delta }(v_k,v_{jk},v_{ijk}))\mathrm{\Delta }_k\}`$
$`+{\displaystyle \underset{i,j,k}{}}ฯต_{ijk}\left\{\mathrm{\Delta }(v_{ik},v_{ijk})\mathrm{\Delta }_{ik}\mathrm{\Delta }(v_{ij},v_{ijk})\mathrm{\Delta }_{ij}\mathrm{\Delta }(v_{jk},v_{ijk})\mathrm{\Delta }_{jk}\right\}`$
$`+{\displaystyle \underset{i,j,k}{}}ฯต_{ijk}v_{ijk}\mathrm{\Delta }_{ijk}.`$
###### Remark A.2.2.
By taking the second augmentation, the total cycle $`\mathrm{\Sigma }`$ maps to:
$$\underset{i,j,k}{}ฯต_{ijk}\mathrm{\Delta }_{ijk},$$
which is the 2-cycle in the CW complex representing the combinatorics of the cover $`๐ฐ`$, and therefore the homology of $`X`$, in degree $`p=2`$.
## Appendix B Acknowledgements
At the early stage of this work we appreciated useful discussions with J.L. Dupont and especially C.-H. Sah, who passed away in July 1997. His generosity of mind and enthusiasm made all our discussions special. He is deeply missed.
The work of L.T. was partially supported by the NSF grant DMS-98-02574.
|
warning/0006/cond-mat0006288.html
|
ar5iv
|
text
|
# Anomalous peak in the superconducting condensate density of cuprate high Tc superconductors at a unique critical doping state
## Abstract
The doping dependence of the superconducting condensate density, n$`{}_{}{}^{o}{}_{s}{}^{}`$, has been studied by muon-spin-rotation for Y<sub>0.8</sub>Ca<sub>0.2</sub>Ba$`{}_{2}{}^{}(`$Cu<sub>1-z</sub>Zn<sub>z</sub>)<sub>3</sub>O<sub>7-ฮด</sub> and Tl<sub>0.5-y</sub>Pb<sub>0.5+y</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Y<sub>x</sub>Cu<sub>2</sub>O<sub>7</sub>. We find that n$`{}_{}{}^{o}{}_{s}{}^{}`$ exhibits a pronounced peak at a unique doping state in the slightly overdoped regime. Its position coincides with the critical doping state where the normal state pseudogap first appears depleting the electronic density of states. A surprising correlation between n$`{}_{}{}^{o}{}_{s}{}^{}`$ and the condensation energy U<sub>o</sub> is observed which suggests unconventional behavior even in the overdoped region.
The superconducting condensate density, n<sub>s</sub>, is proportional to the squared amplitude of the superconducting (SC) order parameter (OP), i.e. of the macroscopic wave function which describes the SC charge carriers. It is thus a fundamental parameter whose variation as a function of temperature (T) and of carrier doping provides important information about the SC state. From early on in the investigation of the cuprate high T<sub>c</sub> superconductors (HTCS), the absolute value of n<sub>s</sub> has been studied by transverse-field muon-spin-rotation (TF-$`\mu `$SR) measurements on polycrystalline samples. Using this technique a linear relationship between the low-T value, n$`{}_{}{}^{o}{}_{s}{}^{},`$ and the critical temperature, T<sub>c</sub>, has been established in underdoped HTCS (so-called โUemura-lineโ) . This finding has stimulated the development of the so-called precursor pairing model whose basic idea is that the low value of n$`{}_{}{}^{o}{}_{s}{}^{}`$ allows for large thermal phase fluctuations of the OP . These phase fluctuations can suppress the formation of a macroscopically coherent SC state over a significant T-interval below the mean-field transition temperature, T$`{}_{}{}^{macr}{}_{c}{}^{}`$$`<`$$`<`$T$`{}_{}{}^{mf}{}_{c}{}^{}`$. This precursor pairing model could also explain the so-called pseudogap effect, which manifests itself as a partial suppression of the low-energy charge and spin excitations in the normal state of underdoped HTCS. In this model the pseudogap state is thought of as the macroscopically incoherent SC state within the range T$`{}_{}{}^{macr}{}_{c}{}^{}`$$`<`$T$`<`$T$`{}_{}{}^{mf}{}_{c}{}^{}`$. Contrasting the precursor pairing model a number of alternative models have been proposed which associate the pseudogap state with electronic correlations which compete with SC . In particular, it has been shown that the combined suppression of T<sub>c</sub> and n$`{}_{}{}^{o}{}_{s}{}^{}`$ in underdoped HTSC can be equally well explained in terms of the depletion of the density of states near the Fermi-level, N($`ฯต_F\pm \mathrm{\Delta }`$), brought about by competing correlations which rapidly grow in strength on the underdoped side . The question as to the origin of the normal state pseudogap is presently vigorously debated and is considered as an important key to resolve the mystery of HTCS.
In this paper we report extensive TF-$`\mu `$SR studies on the evolution of n$`{}_{s}{}^{o}/m_{ab}^{}`$ as a function of hole doping, p, for series of polycrystalline Y<sub>0.8</sub>Ca<sub>0.2</sub>Ba<sub>2</sub>(Cu<sub>1-z</sub>Zn<sub>z</sub>)<sub>3</sub>O<sub>7-ฮด</sub> (Y,Ca-123) and Tl<sub>0.5-y</sub>Pb<sub>0.5+y</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Y<sub>x</sub>Cu<sub>2</sub>O<sub>7</sub> (Tl-1212) samples. Our new data complement previous less detailed studies and demonstrate that n$`{}_{}{}^{o}{}_{s}{}^{}`$ exhibits a pronounced peak at a unique doping state in the slightly overdoped region. Most remarkably, the location of the maximum of n$`{}_{}{}^{o}{}_{s}{}^{}`$ coincides with the critical doping state where previously a rapid suppression of N($`ฯต_F\pm \mathrm{\Delta }`$) has been observed signaling the onset of competing pseudogap correlations . We argue that the sudden and pronounced decrease of n$`{}_{}{}^{o}{}_{s}{}^{}`$ below critical doping cannot readily be explained in terms of the precursor pairing model for which a smooth crossover would be expected. In addition, the strong decrease of n$`{}_{}{}^{o}{}_{s}{}^{}`$ on the overdoped side indicates unconventional behavior even in the absence of the pseudogap correlations.
Series of under- to overdoped polycrystalline samples of Y<sub>0.8</sub>Ca<sub>0.2</sub>Ba<sub>2</sub>(Cu<sub>1-z</sub>Zn<sub>z</sub>)<sub>3</sub>O<sub>7-ฮด</sub> with 0.04$`\delta `$0.98 and z=0, 0.02 and 0.04 and Tl<sub>0.5-y</sub>Pb<sub>0.5+y</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Y<sub>x</sub>Cu<sub>2</sub>O<sub>7</sub> with y$``$0.15 and x$``$0.4 have been prepared following previously described procedures . The T<sub>c</sub> values have been determined by resistivity and DC susceptibility measurements. The hole doping of the CuO<sub>2</sub> planes, p, has been deduced from measurements of the room-temperature thermo-electric power (RT-TEP) . Alternatively, p has been estimated from the ratio of T<sub>c</sub>/T<sub>c,max</sub> (knowing, e.g., from the RT-TEP, whether the sample is under- or overdoped) by assuming the approximate parabolic p-dependence in which p=0.16$`\pm \sqrt{(1T_c/T_{c,\mathrm{max}})/82.6}`$. Good agreement has been obtained between both estimates.
The TF-$`\mu `$SR experiments at an external field of 3 kOe have been performed at the $`\pi `$M3 beamline of the muon facility of the Paul-Scherrer-Institut (PSI) in Villigen/Switzerland. A detailed description of the TF-$`\mu `$SR technique and its use in determining n<sub>s</sub> for polycrystalline HTSC samples is given in reference . A Gaussian relaxation function has been used to fit the measured time spectra. From the obtained Gaussian depolarisation rate, $`\sigma ,`$ we deduced the magnetic penetration depth, $`\lambda _{ab}`$, and the ratio of n<sub>s</sub> to the effective carrier mass m$`{}_{}{}^{}{}_{ab}{}^{}`$ using the established relationship : $`\sigma [\mu s^1]=7.08610^4\lambda _{ab}^2[nm]=2.5110^{21}m_en_s/m_{ab}^{}[cm^3].`$
Figure 1(a) displays the evolution of the low-T value of the depolarisation rate, $`\sigma _o`$n$`{}_{s}{}^{o}/m_{ab}^{},`$ as a function of the hole doping per CuO<sub>2</sub> plane, p, for the series of under- to overdoped Y<sub>0.8</sub>Ca<sub>0.2</sub>Ba<sub>2</sub>(Cu<sub>1-z</sub>Zn<sub>z</sub>)<sub>3</sub>O<sub>7-ฮด</sub> with z=0, 0.02 and 0.04. Figure 1(b) shows the evolution of T<sub>c</sub> with p. It is evident from Fig. 1 that n$`{}_{}{}^{o}{}_{s}{}^{}`$ exhibits a pronounced peak in the slightly overdoped regime. For all three series it occurs at a similar doping state of p $``$ 0.19 (in the following we call it โcritical dopingโ p<sub>crit</sub>). At optimum doping of p$`{}_{opt}{}^{}`$0.16, where the highest T<sub>c</sub> value of T<sub>c,max</sub>=85.5 K is observed for the Zn-free series, n$`{}_{}{}^{o}{}_{s}{}^{}`$ is already reduced by 25-30 % as compared to critical doping. The difference between optimum doping (highest T<sub>c</sub> value) and critical doping (highest n$`{}_{}{}^{o}{}_{s}{}^{}`$ value) is largest for the pure series, but it is reduced for the Zn-substituted series since the optimum doping concentration increases upon Zn-substitution. At very high Zn content, SC survives only in the vicinity of the critical doping state, i.e. p$`{}_{opt}{}^{}`$p<sub>crit</sub>. This effect has been previously explained in terms of the suppression of N($`\epsilon _F\pm \mathrm{\Delta }`$) due to the opening of the pseudogap below critical doping which, for impurity scattering in the unitarity limit, enhances the suppression of T<sub>c</sub> and of n$`{}_{}{}^{o}{}_{s}{}^{}`$ . The other remarkable feature is the plateau in n$`{}_{}{}^{o}{}_{s}{}^{}`$ versus p centered around 1/8 doping. The p-dependence of n$`{}_{}{}^{o}{}_{s}{}^{}`$ therefore is characterized by two marked features, a plateau around 1/8 doping (likely associated with the formation of static stripes) and the peak near critical doping (due to some kind of yet unknown electronic or magnetic correlations which compete with SC). Note that for these Y,Ca-123 samples the contribution of the CuO chains to n$`{}_{}{}^{o}{}_{s}{}^{}`$ should be much weaker than in YBa<sub>2</sub>Cu<sub>4</sub>O<sub>8</sub> (Y-124) or in fully oxygenated Y-123 since the CuO chains are significantly deoxygenated except for the strongly overdoped regime and even there the Ca-substitution leads to chain disorder by occupation of the O(5) off-chain position .
In order to confirm that the peak of n$`{}_{}{}^{o}{}_{s}{}^{}`$ at p$`{}_{crit}{}^{}`$0.19 is a common feature of the hole-doped HTCS we have also investigated a series of under- to overdoped Tl-1212 polycrystalline samples (T<sub>c,max</sub>=107 K). Figure 2 shows the evolution of $`\sigma _o`$n$`{}_{s}{}^{o}/m_{ab}^{}`$ as a function of p. It is evident that $`\sigma _o`$ follows a similar p-dependence as in Y,Ca-123 and, in particular, that it also exhibits a peak at p$`{}_{crit}{}^{}`$0.19.
It is remarkable that the same critical doping state has been previously identified based on specific heat, susceptibility and NMR data as the point where the NS pseudogap first appears and starts to deplete N($`ฯต_F\pm \mathrm{\Delta }`$) . In Fig. 3 the solid line marked by the crosses shows the p-dependence of the ratio of the electronic entropy divided by the temperature, (S/T)$`_{T_c}`$=($`\stackrel{T_c}{}\gamma \left(T^{}\right)dT^{})/T_c,`$ (normalized to the value at critical doping) as obtained by Loram et al. from specific heat data for a similar Y,Ca-123 series . (S/T)$`_{T_c}`$ is the average of $`\gamma `$ between T=0 and T<sub>c</sub> and is a measure of the average density of states within an energy window of $`23`$ k<sub>B</sub>T<sub>c</sub> and thus proportional to N($`ฯต_F\pm \mathrm{\Delta }`$) just above T<sub>c</sub>. It is almost constant on the strongly overdoped side while it exhibits a steady decrease below critical doping due to the opening of the pseudogap. A very similar p-dependence has been obtained from optical measurements for the plasma frequency, $`\omega _{pl}^n`$, of the normal carriers . Shown by the dotted line and marked by the stars is the condensation energy U<sub>o</sub>=$`^{T_c}(S_nS_s)๐T^{}`$ normalized to its value at critical doping . Finally, the normalised condensate density n$`{}_{}{}^{o}{}_{s}{}^{}`$/n$`{}_{}{}^{o}{}_{s,max}{}^{}`$ of the pure Y,Ca-123 and the Tl-1212 series reported here is shown by the open circles and solid triangles, respectively. It is evident from Fig. 3 that n$`{}_{}{}^{o}{}_{s}{}^{}`$ and U<sub>o</sub> follow very similar doping dependencies. Both of them exhibit a pronounced peak around critical doping, they decrease rather steeply on the underdoped as well as on the overdoped sides. Such a correlation is expected on the underdoped side due to the decreasing density of the normal carriers. In fact, (S/T)$`_{T_c}`$ and n$`{}_{}{}^{o}{}_{s}{}^{}`$ can be seen to match very well below p<sub>crit</sub>. Also, the decrease of U<sub>o</sub> on the overdoped side can be understood within a BCS-model due to the decreasing size of T<sub>c</sub> and the concomitant decrease of the size of the SC energy gap (such as observed by spectroscopic techniques ). In clear contrast, the dramatic decrease of n$`{}_{s}{}^{o}/m_{ab}^{}`$ on the overdoped side is completely unexpected since the electronic density of states $``$(S/T)$`_{T_c}`$ or likewise the plasma frequency of the normal carriers $`\omega _{pl}^nn_n/m_{ab}^{}`$ remain almost constant on the overdoped side. We return to this important point later and focus first on the observed behavior on the underdoped side. Below critical doping it is evident that all three quantities which are displayed in Fig. 3 follow a common doping dependence, i.e. below p<sub>crit</sub> they suddenly start to decrease. It has been pointed out earlier that this circumstance of a steady reduction of N($`\epsilon _F\pm \mathrm{\Delta }`$) for p$`<`$p<sub>crit</sub> accompanied by a sharp reduction in the condensation energy U<sub>o</sub> and the condensate density n$`{}_{}{}^{o}{}_{s}{}^{}`$ is precisely what is expected with the onset of an electronic correlation competing with SC . Note that n$`{}_{}{}^{o}{}_{s}{}^{}`$ characterizes the ground state property of a SC. Its sudden change at p<sub>crit</sub> is thus indicative of a drastic alteration of the SC ground-state which is not expected for the precursor pairing model where instead a smooth crossover should occur .
The argument against precursor pairing can be further substantiated by examples which show that T<sub>c</sub> of underdoped samples is not uniquely determined by n$`{}_{}{}^{o}{}_{s}{}^{}`$. YBa<sub>2</sub>Cu<sub>4</sub>O<sub>8</sub> (Y-124) has, besides the superconducting CuO<sub>2</sub> bilayers, metallic CuO chains that become superconducting most likely due to proximity coupling . Y-124 is underdoped with T<sub>c</sub>=80 K and exhibits clear signatures of the pseudogap. The metallic CuO double chains, however, lead to a significant enhancement of the condensate density (with $`\sigma _o`$=3.3 $`\mu `$s$`{}_{s}{}^{1})`$ as compared to similarly underdoped Y-123 or Y,Ca-123 with T$`{}_{c}{}^{}`$80 K and $`\sigma _o`$2.2-2.3 $`\mu `$s<sup>-1</sup> . If T<sub>c</sub> of underdoped samples was indeed uniquely determined by n$`{}_{}{}^{o}{}_{s}{}^{}`$, then T<sub>c</sub> should exceed 100 K in Y-124. Another counter example is underdoped Zn-substituted Y-123. It has been shown that upon Zn-substitution n$`{}_{}{}^{o}{}_{s}{}^{}`$ is even more rapidly suppressed than T<sub>c</sub>. This behavior has been explained in terms of the d-wave symmetry of the superconducting OP and elastic scattering in the unitarity limit on Zn-impurities . In a plot of T<sub>c</sub> versus n$`{}_{}{}^{o}{}_{s}{}^{}`$ such compounds thus lie far to the left of the โUemura lineโ. For Y<sub>0.8</sub>Ca<sub>0.2</sub>Ba<sub>2</sub>Cu<sub>2.94</sub>Zn<sub>0.06</sub>O<sub>6.45</sub> with T$`{}_{c}{}^{}=`$31 K we obtained $`\sigma _o`$0.2 $`\mu `$s$`{}_{}{}^{1}{}_{s}{}^{}`$ which according to the Uemura-relation should result in T$``$10 K, i.e. one third the observed value . Finally, there is the surprising result that n$`{}_{}{}^{o}{}_{s}{}^{}`$ decreases very strongly also on the overdoped side . Despite the small n$`{}_{}{}^{o}{}_{s}{}^{}`$ values no pseudogap effect is observed for strongly overdoped samples. These examples imply that it is not primarily the small value of n$`{}_{}{}^{o}{}_{s}{}^{}`$ which is responsible for the low T<sub>c</sub> values of underdoped HTCS or which causes the pseudogap effect. This conclusion is in clear contrast to the precursor pairing model.
Finally, we focus on the TF-$`\mu `$SR data on the strongly overdoped side past critical doping. Our measurements confirm previous reports that n$`{}_{s}{}^{o}/m_{ab}^{}`$ is dramatically reduced on overdoping. This very surprising result was first obtained by TF-$`\mu `$SR on Tl<sub>2</sub>Ba<sub>2</sub>CuO<sub>6+ฮด</sub> (Tl-2201) and later on (Yb,Ca)-123 and (Y,Ca)-123 . Subsequently, it has been confirmed by other experimental techniques . It was previously pointed out that the strong suppression of n$`{}_{s}{}^{o}/m_{ab}^{}`$ for heavily overdoped samples cannot be understood within a BCS model, unless one assumes that pairbreaking correlations become increasingly important . This earlier proposal of strong pair breaking on the overdoped side, however, is not supported by recent specific heat or <sup>89</sup>Y- and <sup>17</sup>O-NMR data which give no clear indication for a growing density of unpaired carriers within the SC gap. These data also do not support the scenario that overdoped materials are inhomogeneous with only a small SC fraction . As was mentioned above, the dramatic decrease of n$`{}_{s}{}^{o}/m_{ab}^{}`$ in the overdoped region is completely unexpected. The BCS-model predicts that n$`{}_{}{}^{o}{}_{s}{}^{}`$/m$`{}_{}{}^{}{}_{ab}{}^{}`$ should follow the same p-dependence as the normal state plasma frequency, $`\omega _{pl}^nn_n/m_{ab}^{},`$ deduced from optical experiments or (S/T)$`_{T_c}`$ obtained from the specific heat which both remain almost constant on the overdoped side. In this context we would like to emphasize the surprising similarity between the p-dependences of n$`{}_{}{}^{o}{}_{s}{}^{}`$ as deduced from TF-$`\mu `$SR and the condensation energy U<sub>o</sub> obtained from the specific heat measurements . Figure 3 implies that $`n_s^o`$ and U<sub>o</sub> are correlated over the entire doping range. Such a relationship cannot easily be understood within BCS theory where the condensation energy U<sub>o</sub> is determined by the change in potential energy due to the attractive pairing interaction times the density of state N($`\epsilon _F\pm \mathrm{\Delta }`$), while n$`{}_{}{}^{s}{}_{o}{}^{}`$ should be determined by the kinetic energy of the carriers times N($`\epsilon _F\pm \mathrm{\Delta }`$). This apparent inconsistency of the experimental data on the overdoped side with the prediction of the BCS-model is especially important in the light of the indication that competing electronic correlations are at work below critical doping. This should mean that the intrinsic properties of the SC state are best seen for overdoped materials. Yet it is this very region which cannot be described by the BCS-model and therefore suggest an unconventional SC state. As one example of an unconventional model which explains the unusual correlation between U<sub>o</sub> and n$`{}_{}{}^{o}{}_{s}{}^{}`$ we only mention here the so-called spin-charge separation model. Superconductivity can only occur here if both holons and spinons condense resulting in a total condensate density of 1/n<sub>s</sub>=1/n$`{}_{}{}^{holon}{}_{s}{}^{}`$+1/n$`{}_{}{}^{spinon}{}_{s}{}^{}`$ . On the overdoped side the diminishing spinon density would therefore lead to the dramatic reduction of n$`{}_{}{}^{o}{}_{s}{}^{}`$. As a final remark we note that a similar TF-$`\mu `$SR study has recently been performed on the p-dependence of n$`{}_{}{}^{o}{}_{s}{}^{}`$ in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> (La-214) . For La-214 the maximum in n$`{}_{}{}^{o}{}_{s}{}^{}`$ versus p is significantly broader than the one reported here for Y,Ca-123 and Tl-1212. The reason for this probably lies in the fact that N($`\epsilon _F\mathrm{\Delta }`$) for La-214 goes on increasing with overdoping . What is important is that while n$`{}_{}{}^{o}{}_{s}{}^{}`$/(S/T)$`_{T_c}`$ for La-214 remains constant below p<sub>crit</sub>, this ratio decreases sharply for p$`>`$p<sub>crit</sub> just as shown in Fig. 3 for Y,Ca-123 and Tl-1212. Moreover, the peak in U<sub>o</sub> for La-214 is also significantly broadened as compared with the one in Y,Ca-123 . This finding seems to support our suggestion that n$`{}_{}{}^{o}{}_{s}{}^{}`$ and U<sub>o</sub> may be correlated in the HTSC.
In summary, we have presented experimental evidence based on TF-$`\mu `$SR measurements that a unique critical doping state exists in the slightly overdoped regime where the superconducting condensate density, n$`{}_{}{}^{o}{}_{s}{}^{}`$, exhibits a pronounced maximum. The observed sudden change of n$`{}_{}{}^{o}{}_{s}{}^{}`$ is not expected within the precursor pairing model . On the underdoped side the decrease of n$`{}_{}{}^{o}{}_{s}{}^{}`$ coincides with the decrease of the normal density of states caused by the pseudogap correlations. This effect sets in abruptly at critical doping . In contrast, the rapid decrease of n$`{}_{}{}^{o}{}_{s}{}^{}`$ on the strongly overdoped side cannot be easily understood within a conventional BCS model since the normal state carrier density remains almost constant here. Even more surprisingly, n$`{}_{}{}^{o}{}_{s}{}^{}`$ seems to scale with the condensation energy U<sub>o</sub> as obtained from specific heat measurements . This correlation suggests unconventional behavior even in the overdoped region.
We acknowledge A. Amato and D. Herlach (PSI) for technical support. C.B. and J.L.T thank J.W. Loram for fruitful discussions and for providing the specific heat data. C.B. appreciates financial support of the Marsden Fund of New Zealand during his stay at IRL. J.L.T. thanks the Royal Society of New Zealand for financial support under a James Cook Fellowship.
Figure Captions
Figure 1: (a) Evolution of the low-T depolarisation rate $`\sigma _o`$n$`{}_{s}{}^{o}/m_{ab}^{}`$ as a function of hole doping, p, for series of under- to overdoped Y<sub>0.8</sub>Ca<sub>0.2</sub>Ba<sub>2</sub>Cu<sub>3-z</sub>Zn<sub>z</sub>O<sub>7-ฮด</sub> with z=0 (open circles), z=0.02 (solid squares) and z=0.04 (stars). The crossed circle shows $`\sigma _o`$ for strongly overdoped Lu<sub>0.7</sub>Ca<sub>0.3</sub>Ba<sub>2</sub>Cu<sub>3</sub>O<sub>6.95</sub>. Critical (optimum) doping is marked by the solid (dotted) line. (b) Doping dependence of the critical temperature T<sub>c</sub> shown by the same symbols.
Figure 2: Doping dependence of the low-T depolarisation rate $`\sigma _o`$ for under- to overdoped Tl<sub>1-y</sub>Pb<sub>y</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Y<sub>x</sub>Cu<sub>2</sub>O<sub>7</sub>. The solid (dotted) line marks critical (optimum) doping.
Figure 3: Doping dependence of the normalized values of (S/T)$`_{T_c}`$ (solid line and crosses) and of U<sub>o</sub> (solid line and stars) of Y,Ca-123 as deduced from specific heat and of n$`{}_{}{}^{o}{}_{s}{}^{}`$/m$`{}_{}{}^{}{}_{ab}{}^{}`$ for Y,Ca-123 (open circles) and Tl-1212 (solid triangles) deduced from TF-$`\mu `$SR.
|
warning/0006/hep-th0006116.html
|
ar5iv
|
text
|
# Reparametrization-Invariant Path Integral in GR and โBig Bangโ of Quantum Universe
## 1. Introduction
โBig Bangโ as the beginning of the Hubble evolution of a universe is described as a pure classical phenomenon on the basic of particular solutions of the Einstein equations in general relativity in the homogeneous approximation. A strange situation consists in that the highest level of the theory, i.e., the Faddeev-Popov-DeWitt generating functional for unitary S-matrix , neglects the questions about the evolution of a universe which are in the competence of the simplest classical approximation. There is an opinion that the solution of the โBig Bangโ problem in quantum theory goes beyond the scope of the unitary perturbation theory and even of general relativity. To answer these questions, we need a more general theory of the type of superstring .
According to another point of view, the reason of theoretical difficulties in understanding the โBig Bangโ phenomenon is not the Einstein theory, but the non-invariant method of its quantization. In particular, for the Faddeev-Popov-DeWitt unitary S-matrix the non-invariant coordinate time is considered as the time of evolution, whereas an observer in a universe can observe and measure only invariants of group of diffeomorphisms of the Hamiltonian dynamics, which includes reparametrizations of the coordinate time .
In the present paper we try to construct the unitary S-matrix for general relativity in a finite space-time in terms of the reparametrization-invariant evolution parameter, and to answer the questions: What do Quantum Universe and Quantum Gravity mean? What is the status of the โBig Bangโ evolution in quantum theory? What does creation of Quantum Universe mean? on the level of perturbation theory, using the scheme of the time-reparametrization-invariant Hamiltonian reduction .
In the context of the Dirac generalized Hamiltonian theory for constrained systems , this scheme means the explicit resolving of the first class constraints to determine the constraint-shell action directly in terms of invariants. In other words, we use the invariant reduction of the action (to obtain an equivalent unconstrained system) instead of the generally accepted non-invariant reduction of the phase space by fixing gauges (see Fig. 1).
The example of the application of such an invariant reduction of the action is the Dirac formulation of QED directly in terms of the gauge-invariant (dressed) fields as the proof of the adequateness of the Coulomb gauge with the invariant content of classical equations. Recall that the invariant reduction of the action is the way to obtain the unconstrained Feynman integral for the foundation of the intuitive Faddeev-Popov functional integral in gauges theories and to reveal collective excitations of the gauge fields in the form of zero-modes of the first class constraints (see Fig. 2).
A constructive idea of the considered invariant Hamiltonian reduction of general relativity is to introduce the dynamic evolution parameter as the zero-mode collective excitation of metric . This dynamic evolution parameter can be identified with the zero-Fourier harmonic of the space-metric determinant (treated in cosmology as the cosmic scale factor), whereas its conjugate momentum, i.e, the second (external) form, plays the role of the localizable Hamiltonian of evolution.
The separation of this zero-mode evolution parameter on the level of the action allows us to determine also the invariant geometric time formed by averaging the time-like component of a metric over the space coordinates .
An observer reveals the evolution of the universe (with โBig Bangโ) as the dependence of the dynamic evolution parameter (i.e. cosmic scale factor) on the geometric time.
The evolution of both a classical and a quantum universes in terms of the geometric time has the form of the canonical transformations of the initial dynamic variables to a new set of variables for which the total energy constraint becomes a new momentum, and its conjugate variable (i.e., a new dynamic evolution parameter) coincides with the geometric time onto equations of motion.
The content of the paper is the following. Section 2 is devoted to the reparametrization-invariant Hamiltonian reduction of general relativity. In section 3 we construct the generating functional for the unitary perturbation theory which includes โBig Bangโ and Hubble evolution. In Section 4, โBig Bangโ and Hubble evolution are reproduced in lowest order of perturbation theory. In Section 5, we research the conditions of validity of the conventional quantum field theory in the infinite space-time limit.
## 2. Reparametrization-invariant Hamiltonian dynamics of GR
### 2.1. Action and variables
General relativity (GR) is given by the singular Einstein-Hilbert action with the matter fields
$$W(g|\mu )=d^4x\sqrt{g}[\frac{\mu ^2}{6}R(g)+_f]\left(\mu ^2=M_{Planck}^2\frac{3}{8\pi }\right)$$
(1)
and by a measurable interval
$$(ds)_e^2=g_{\alpha \beta }dx^\alpha dx^\beta .$$
(2)
They are invariant with respect to general coordinate transformations
$$x_\mu x_\mu ^{}=x_\mu ^{}(x_0,x_1,x_2,x_3).$$
(3)
The generalized Hamiltonian approach to GR was formulated by Dirac and Arnovit, Deser and Misner as a theory of system with constraints in $`3+1`$ foliated space-time
$$(ds)^2=g_{\mu \nu }dx^\mu dx^\nu =N^2dt^2{}_{}{}^{(3)}g_{ij}^{}\stackrel{ห}{dx}{}_{}{}^{i}\stackrel{ห}{dx}{}_{}{}^{j}(\stackrel{ห}{dx}{}_{}{}^{i}=dx^i+N^idt)$$
(4)
with the lapse function $`N(t,\stackrel{}{x})`$, three shift vectors $`N^i(t,\stackrel{}{x})`$, and six space components $`{}_{}{}^{(3)}g_{ij}^{}(t,\stackrel{}{x})`$ depending on the coordinate time $`t`$ and the space coordinates $`\stackrel{}{x}`$. The Dirac-ADM parametrization of metric (4) characterizes a family of hypersurfaces $`t=\mathrm{const}.`$ with the unit normal vector $`\nu ^\alpha =(1/N,N^k/N)`$ to a hypersurface and with the second (external) form
$$\frac{1}{N}({}_{}{}^{(3)}\dot{g}_{ij}^{})\mathrm{\Delta }_iN_j\mathrm{\Delta }_jN_i)$$
(5)
that shows how this hypersurface is embedded into the four-dimensional space-time.
Coordinate transformations conserving the family of hypersurfaces $`t=\mathrm{const}.`$
$$t\stackrel{~}{t}=\stackrel{~}{t}(t);x_i\stackrel{~}{x}_i=\stackrel{~}{x}_i(t,x_1,x_2,x_3),$$
(6)
$$\stackrel{~}{N}=N\frac{dt}{d\stackrel{~}{t}};\stackrel{~}{N}^k=N^i\frac{\stackrel{~}{x}^k}{x_i}\frac{dt}{d\stackrel{~}{t}}\frac{\stackrel{~}{x}^k}{x_i}\frac{x^i}{\stackrel{~}{t}}$$
(7)
are called a kinemetric subgroup of the group of general coordinate transformations (3. The group of kinemetric transformations is the group of diffeomorphisms of the generalized Hamiltonian dynamics. It includes reparametrizations of the nonobservable time coordinate $`\stackrel{~}{t}(t)`$) (6) that play the principal role in the procedure of the reparametrization-invariant reduction discussed in the previous Sections. The main assertion of the invariant reduction is the following: the dynamic evolution parameter is not the coordinate but the variable with a negative contribution to the energy constraint. (Recall that this reduction is based on the explicit resolving of the global energy constraint with respect to the conjugate momentum of the dynamic evolution parameter to convert this momentum into the Hamiltonian of evolution of the reduced system.)
A negative contribution to the energy constraint is given by the space-metric-determinant logarithm. Therefore, following papers we introduce an invariant evolution parameter $`\phi _0(t)`$ as the zero Fourier harmonic component of this logarithm (treated, in cosmology, as the cosmic scale factor). This variable is distinguished in general relativity by the Lichnerowicz conformal-type transformation of field variables $`f`$ with the conformal weight $`(n)`$
$${}_{}{}^{(n)}\overline{f}={}_{}{}^{(n)}f\left(\frac{\phi _0(t)}{\mu }\right)^n,$$
(8)
where $`n=2,\mathrm{\hspace{0.33em}0},3/2,1`$ for the tensor, vector, spinor, and scalar fields, respectively, $`\overline{f}`$ is so-called conformal-invariant variable used in GR for the analysis of initial data . In particular, for metric we get
$$g_{\mu \nu }(t,\stackrel{}{x})=\left(\frac{\phi _0(t)}{\mu }\right)^2\overline{g}_{\mu \nu }(t,\stackrel{}{x})(ds)^2=\left(\frac{\phi _0(t)}{\mu }\right)^2[\overline{N}^2dt^2{}_{}{}^{(3)}\overline{g}_{ij}^{}\stackrel{ห}{dx}{}_{}{}^{i}\stackrel{ห}{dx}{}_{}{}^{j}].$$
(9)
As the zero Fourier harmonic is extracted from the space metric determinant logarithm, the space metric $`\overline{g}_{ij}(t,\stackrel{}{x})`$ should be defined in a class of nonzero harmonics
$$d^3x\mathrm{log}\overline{g}_{ij}(t,\stackrel{}{x})=0.$$
(10)
The transformational properties of the curvature $`R(g)`$ with respect to the transformations (9) lead to the action (1) in the form
$$W(g|\mu )=W(\overline{g}|\phi _0)\underset{t_1}{\overset{t_2}{}}๐t\underset{V_0}{}d^3x\phi _0\frac{d}{dt}(\frac{\dot{\phi }_0\sqrt{\overline{g}}}{\overline{N}}).$$
(11)
This form define the global lapse function $`N_0`$ as the average of the lapse function $`\overline{N}`$ in the metric $`\overline{g}`$ over the kinemetric invariant space volume
$$N_0(t)=\frac{V_0}{\underset{V_0}{}d^3x\frac{\sqrt{\overline{g}(t,\stackrel{}{x}}}{\overline{N}(t,\stackrel{}{x})}},\overline{g}=det({}_{}{}^{(3)}\overline{g}),V_0=\underset{V_0}{}d^3x,$$
(12)
where $`V_0`$ is a free parameter which in the perturbation theory has the meaning of a finite volume of the free coordinate space. The lapse function $`\overline{N}(t,\stackrel{}{x})`$ can be factorized into the global component $`N_0(t)`$ and the local one $`๐ฉ(t,\stackrel{}{x})`$
$$\overline{N}(t,\stackrel{}{x})\overline{g}^{1/2}:=N_0(t)๐ฉ(t,\stackrel{}{x}):=N_q,$$
(13)
where $`๐ฉ`$ fulfills normalization condition:
$$I[๐ฉ]:=\frac{1}{V_0}\frac{d^3x}{๐ฉ}=1$$
(14)
that is imposed after the procedure of variation of action, to reproduce equations of motion of the initial theory. In the Dirac harmonical variables chosen as
$$q^{ik}=\overline{g}\overline{g}^{ik},$$
(15)
the metric (4) takes the form
$$(ds)^2=\frac{\phi _0(t)^2}{\mu ^2}q^{1/2}(N_q^2dt^2q_{ij}\stackrel{ห}{dx}{}_{}{}^{i}\stackrel{ห}{dx}{}_{}{}^{j}),(q=det(q^{ij})).$$
(16)
The Dirac-Bergmann version of action (11) in terms of the introduced above variables reads
$$W=\underset{t_1}{\overset{t_2}{}}๐t\left\{L+\frac{1}{2}_t(P_0\phi _0)\right\},$$
(17)
$$L=\left[\underset{V_0}{}d^3x\left(\underset{F}{}P_F\dot{F}N^i๐ซ_i\right)\right]P_0\dot{\phi }_0N_0\left[\frac{P_0^2}{4V_0}+I^1H(\phi _0)\right],$$
(18)
where
$$\underset{F}{}P_F\dot{F}=\underset{f}{}p_f\dot{f}\pi _{ij}\dot{q}^{ij};$$
(19)
$$H(\phi _0)=d^3x๐ฉ(\phi _0)$$
(20)
is the total Hamiltonian of the local degrees of freedom,
$$(\phi _0)=\frac{6}{\phi _0^2}q^{ij}q^{kl}[\pi _{ik}\pi _{jl}\pi _{ij}\pi _{kl}]+\frac{\phi _0^2q^{1/2}}{6}{}_{}{}^{(3)}R(\overline{g})+_f,$$
(21)
and
$$๐ซ_i=2[_k(q^{kl}\pi _{il})_i(q^{kl}\pi _{kl})]+๐ซ_{if}$$
(22)
are the densities of energy and momentum and $`_f,๐ซ_f`$ are contributions of the matter fields. In the following, we call the set of the field variables $`F`$ (19) with the dynamic evolution parameter $`\phi _0`$ the field world space. The local part of the momentum of the space metric determinant
$$\pi (t,x):=q^{ij}\pi _{ij}$$
(23)
is given in the class of functions with the non-zero Fourier harmonics, so that
$$d^3x\pi (t,x)=0.$$
(24)
The geometric foundation of introducting the global variable (9) in GR was given in as the assertion about the nonzero value of the second form in the whole space. This assertion (which contradicts the Dirac gauge $`\pi =0`$) follows from the global energy constraint, as, in the lowest order of the Dirac perturbation theory, positive contributions of particle-like excitations to the zero Fourier harmonic of the energy constraint can be compensated only by the nonzero value of the second form.
The aim of this paper is to obtain the dynamic โequivalentโ unconstrained system in the field world space ($`F,\phi _0`$) by explicit resolving the global energy constraint and to study the possibility of the Hamiltonian description of the โequivalentโ unconstrained system in terms of the reparametrization-invariant evolution parameter $`T`$ defined by equation $`N_0dt=dT`$.
### 2.2. Local constraints and equations of motion
Following Dirac we formulate generalized Hamiltonian dynamics for the considered system (17). It means the inclusion of momenta for $`๐ฉ`$ and $`N_i`$ and appropriate terms with Lagrange multipliers
$$W^D=\underset{t_1}{\overset{t_2}{}}๐t\left\{L^D+\frac{1}{2}_t(P_0\phi _0)\right\},L^D=L+d^3x(P_๐ฉ\dot{๐ฉ}+P_{N^i}\dot{N}^i\lambda ^0P_๐ฉ\lambda ^iP_{N^i}).$$
(25)
We can define extended Dirac Hamiltonian as
$$H^D=N_0\left[\frac{P_0^2}{4V_0}+I^1H(\phi _0)\right]+d^3x(\lambda ^0P_๐ฉ+\lambda ^iP_{N^i}).$$
(26)
The equations obtained from variation of $`W^D`$ with respect to Lagrange multipliers are called first class primary constraints
$$P_๐ฉ=0,P_{N^i}=0.$$
(27)
The condition of conservation of these constraints in time leads to the first class secondary constraints
$$\{H^D,P_๐ฉ\}=\frac{d^3x๐ฉ}{V_0๐ฉ^2}=0,\{H^D,P_{N^i}\}=๐ซ_i=0.$$
(28)
For completeness of the system we have to include set of secondary constraints. According Dirac we choose them in the form
$$๐ฉ(t,\stackrel{}{x})=1;N^i(t,\stackrel{}{x})=0;$$
(29)
$$\pi (t,\stackrel{}{x})=0;\chi ^j:=_i(q^{1/3}q^{ij})=0.$$
(30)
The equations of motion obtained for our system are
$$\frac{dF}{dT}=\frac{H(\phi _0)}{P_F},\frac{dP_F}{dT}=\frac{H(\phi _0)}{F},$$
(31)
where $`H(\phi _0)`$ is given by the equation (20), and we introduced reparametrization-invariant geometric time $`T`$
$$N_0dt\stackrel{\mathrm{def}}{=}\mathrm{dT}.$$
(32)
### 2.3. Global constraints and equations of motion.
The physical meaning of the geometric time $`T`$, the dynamic variable $`\phi _0`$ and its momentum is given by the explicit resolving of the zero-Fourier harmonic of the energy constraint
$$\frac{\delta W^E}{\delta N_0(t)}=\frac{P_0^2}{4V_0}+H(\phi _0)=0.$$
(33)
This constraint has two solutions for the global momentum $`P_0`$:
$$(P_0)_\pm =\pm 2\sqrt{V_0H(\phi _0)}H_\pm ^{}.$$
(34)
The equation of motion for this global momentum $`P_0`$ in gauge (29) takes the form
$$\frac{\delta W^E}{\delta P_0}=0\left(\frac{d\phi }{dT}\right)_\pm =\frac{(P_0)_\pm }{2V}=\pm \sqrt{\rho (\phi _0)};\rho (\phi _0)=\frac{d^3x}{V_0}=\frac{H(\phi _0)}{V_0}.$$
(35)
The integral form of the last equation is
$$T(\phi _1,\phi _2)=\underset{\phi _1}{\overset{\phi _2}{}}๐\phi \rho ^{1/2}(\phi ).$$
(36)
Equation obtained by varying the action with respect to $`\phi _0`$ follows independently from the set of other constraints and equations of motion.
Equations (35), (36) in the homogeneous approximation of GR are the basis of observational cosmology where the geometric time is the conformal time connected with the world time $`T_f`$ of the Friedmann cosmology by the relation
$$dT_f=\frac{\phi _0(T)}{\mu }dT,$$
(37)
and the dependence of scale factor (dynamic evolution parameter $`\phi _0`$) on the geometric time $`T`$ is treated as the evolution of the universe. In particular, equation (35) gives the relation between the present-day value of the dynamic evolution parameter $`\phi _0(T_0)`$ and cosmological observations, i.e., the density of matter $`\rho `$ and the Hubble parameter
$$_{hub}^e=\frac{\mu \phi _0^{}}{\phi _0^2}=\frac{\mu \sqrt{\rho }}{\phi _0^2}\phi _0(T_0)=\left(\frac{\mu \sqrt{\rho }}{_{hub}}\right)^{1/2}:=\mu \mathrm{\Omega }_0^{1/4}(0.6<(\mathrm{\Omega }_0^{1/4})_{exp}<1.2).$$
(38)
The dynamic evolution parameter as the cosmic scale factor and a value of its conjugate momentum (i.e., a value of the dynamic Hamiltonian) as the density of matter (see equations (35), (36)) are objects of measurement in observational astrophysics and cosmology and numerous discussions about the Hubble parameter, dark matter, and hidden mass.
Our aim is to find the equivalent unconstrained Hamiltonian system that describes evolution of the field world space $`(F,\phi _0)`$ in terms of the geometric time $`T`$.
### 2.4. Equivalent Unconstrained Systems
Assume that we can solve the constraint equations and pass to the reduced space of independent variables ($`F^{},P_F^{}`$). The explicit solution of the local and global constraints has two analytic branches with positive and negative values for scale factor momentum $`P_0`$ (34). Therefore, inserting solutions of all constraints into the action we get two branches of the equivalent Dynamic Unconstrained System (DUS)
$$W_\pm ^{DUS}[F|\phi _0]=\underset{\phi _1}{\overset{\phi _2}{}}๐\phi _0\left\{\left[d^3x\underset{F^{}}{}P_F^{}\frac{F^{}}{\phi _0}\right]H_\pm ^{}+\frac{1}{2}_{\phi _0}(\phi _0H_\pm ^{})\right\},$$
(39)
where $`\phi _0`$ plays the role of evolution parameter and $`H_\pm ^{}`$ (defined by equation (34) plays the role of the evolution Hamiltonian in the reduced phase space of independent physical variables $`(F^{},P_F^{})`$ with equations of motion
$$\frac{dF^{}}{d\phi _0}=\frac{H_\pm ^{}}{P_F^{}},\frac{dP_F^{}}{d\phi _0}=\frac{H_\pm ^{}}{F^{}}.$$
(40)
The evolution of the field world space variables $`(F^{},\phi _0)`$ with respect to the geometric time $`T`$ is not contained in DUS (39). This geometric time evolution is described by supplementary equation for nonphysical momentum $`P_0`$ (35) that follows from the initial extended system.
To get an equivalent unconstrained system in terms of the geometric time (we call it the Geometric Unconstrained System (GUS)), we need the Levi-Civita canonical transformation (LC) of the field world phase space
$$(F^{},P_F^{}|\phi _0,P_0)(F_G^{},P_G^{}|Q_0,\mathrm{\Pi }_0)$$
(41)
which converts the energy constraint (33) into the new momentum $`\mathrm{\Pi }_0`$ (see the similar transformations for a relativistic particle in Appendix A).
In terms of geometrical variables the action takes the form
$$W^G=\underset{t_1}{\overset{t_2}{}}๐t\left\{\left[d^3x\underset{F_G^{}}{}P_G^{}\dot{F}_G^{}\right]\mathrm{\Pi }_0\dot{Q}_0+N_0\mathrm{\Pi }_0+\frac{d}{dt}S^{LC}\right\}$$
(42)
where $`S^{LC}`$ is generating function of LC transformations. Then the energy constraint and the supplementary equation for the new momentum take trivial form
$$\mathrm{\Pi }_0=0;\frac{\delta W}{\delta \mathrm{\Pi }_0}=0\frac{dQ_0}{dt}=N_0dQ_0=dT.$$
(43)
Equations of motion are also trivial
$$\frac{dP_G^{}}{dT}=0,\frac{dF_G^{}}{dT}=0,$$
(44)
and their solutions are given by the initial data
$$P_G^{}=P_{G}^{}{}_{}{}^{0},F_G^{}=F_{G}^{}{}_{}{}^{0}.$$
(45)
Substituting solutions of (43) and (44) into the inverted Levi-Civita transformations
$$F^{}=F^{}(Q_0,\mathrm{\Pi }_0|F_G^{},P_G^{}),\phi _0=\phi _0(Q_0,\mathrm{\Pi }_0|F_G^{},P_G^{})$$
(46)
and similar for momenta, we get formal solutions of (40) and (36)
$$F^{}=F^{}(T,0|F_{G}^{}{}_{}{}^{0},P_{G}^{}{}_{}{}^{0}),P_F^{}=P_F^{}(T,0|F_{G}^{}{}_{}{}^{0},P_{G}^{}{}_{}{}^{0}),\phi _0=\phi _0(T,0|F_{G}^{}{}_{}{}^{0},P_{G}^{}{}_{}{}^{0}).$$
(47)
We see that the geometric time evolution of the dynamic variables is absent in DUS. The geometric time evolution of the dynamic variables can be described in the form of the LC (inverted) canonical transformation of GUS into DUS (46), (47).
To obtain the geometric time Hamiltonian evolution, it is sufficient to use the weak form of Levi-Civita-type transformations to GUS $`(F^{},P_F^{})(\stackrel{~}{F},\stackrel{~}{P})`$ with a new constraint
$$\stackrel{~}{\mathrm{\Pi }}_0\stackrel{~}{H}(\stackrel{~}{Q}_0,\stackrel{~}{F},\stackrel{~}{P})=0.$$
(48)
We get the constraint shell action
$$\stackrel{~}{W}^{GUS}=๐T\left\{\left[d^3x\underset{\stackrel{~}{F}}{}\stackrel{~}{P}\frac{d\stackrel{~}{F}}{dT}\right]\stackrel{~}{H}(T,\stackrel{~}{F},\stackrel{~}{P})\right\},$$
(49)
that allows us to choose the initial cosmological data with respect to the geometric time.
The considered invariant reduction reveals the difference of reparametrization-invariant theory from the gauge-invariant theory: in gauge-invariant theory the superfluous (longitudinal) variables are completely excluded from the reduced system; whereas, in reparametrization-invariant theory the superfluous (longitudinal) variables leave the sector of the Dirac observables (i.e., the phase space ($`F^{},P_F^{}`$)) but not the sector of measurable quantities: superfluous (longitudinal) variables become the dynamic evolution parameter and dynamic Hamiltonian of the reduced theory.
### 2.5. Quantization and the arrow of the time
In quantum theory of GR (like in quantum theories of a particle considered in Appendix A), we get two Schrรถdinger equations
$$i\frac{d}{d\phi _0}\mathrm{\Psi }^\pm (F|\phi _0,\phi _1)=H_\pm ^{}(\phi _0)\mathrm{\Psi }^\pm (F|\phi _0,\phi _1)$$
(50)
with positive and negative eigenvalues of $`P_0`$ and normalizable wave functions with the spectral series over quantum numbers $`Q`$
$$\mathrm{\Psi }^+(F|\phi _0,\phi _1)=\underset{Q}{}A_Q^+<F|Q><Q|\phi _0,\phi _1>\theta (\phi _0\phi _1)$$
(51)
$$\mathrm{\Psi }^{}(F|\phi _0,\phi _1)=\underset{Q}{}A_Q^{}<F|Q>^{}<Q|\phi _0,\phi _1>^{}\theta (\phi _1\phi _0),$$
(52)
where $`<F|Q>`$ is the eigenfunction of the reduced energy (34)
$$H_\pm ^{}(\phi _0)<F|Q>=\pm E(Q,\phi _0)<F|Q>$$
(53)
$$<Q|\phi _0,\phi _1>=\mathrm{exp}[i\underset{\phi _1}{\overset{\phi _0}{}}๐\phi E(Q,\phi )],<Q|\phi _0,\phi _1>^{}=\mathrm{exp}[i\underset{\phi _1}{\overset{\phi _0}{}}๐\phi E(Q,\phi )].$$
(54)
The coefficient $`A_Q^+`$, in โsecondaryโ quantization, can be treated as the operator of creation of a universe with positive energy; and the coefficient $`A_Q^{}`$, as the operator of annihilation of a universe also with positive energy. The โsecondaryโ quantization means $`[A_Q^{},A_Q^{}^+]=\delta _{Q,Q^{}}`$. The physical states of a quantum universe are formed by the action of these operators on the vacuum $`<0|`$, $`|0>`$ in the form of out-state ($`|Q>=A_Q^+|0>`$) with positive โfrequenciesโ and in-state ($`<Q|=<0|A_Q^{}`$) with negative โfrequenciesโ. This treatment means that positive frequencies propagate forward ($`\phi _0>\phi _1`$); and negative frequencies, backward ($`\phi _1>\phi _0`$), so that the negative values of energy are excluded from the spectrum to provide the stability of the quantum system in quantum theory of GR. In other words, instead of changing the sign of energy, we change that of the dynamic evolution parameter, which leads to the causal Green function
$$G_c(F_1,\phi _1|F_2,\phi _2)=G_+(F_1,\phi _1|F_2,\phi _2)\theta (\phi _2\phi _1)+G_{}(F_1,\phi _1|F_2,\phi _2)\theta (\phi _1\phi _2)$$
(55)
where $`G_+(F_1,\phi _1|F_2,\phi _2)=G_{}(F_2,\phi _2|F_1,\phi _1)`$ is the โcommutativeโ Green function
$$G_+(F_2,\phi _2|F_1,\phi _1)=<0|\mathrm{\Psi }^{}(F_2|\phi _2,\phi _1)\mathrm{\Psi }^+(F_1|\phi _1,\phi _1)|0>$$
(56)
For this causal convention, the geometric time (36) is always positive in accordance with the equations of motion (35)
$$\left(\frac{dT}{d\phi _0}\right)_\pm =\pm \sqrt{\rho }T_\pm (\phi _1,\phi _0)=\pm \underset{\phi _1}{\overset{\phi _0}{}}๐\phi \rho ^{1/2}(\phi )0.$$
(57)
Thus, the causal structure of the field world space immediatly leads to the arrow of the geometric time (57) and the beginning of evolution of a universe with respect to the geometric time $`T=0`$.
As it was shown in , the way to obtain conserved integrals of motion in classical theory and quantum numbers $`Q`$ in quantum theory is the Levi-Civita-type canonical transformation of the field world space $`(F,\phi _0)`$ to a geometric set of variables ($`V,Q_0`$) with the condition that the geometric evolution parameter $`Q_0`$ coincides with the geometric time $`dT=dQ_0`$ (see Fig. 3).
## 3. Reparametrization invariant path integral
Following Faddeev-Popov procedure we can write down the path integral for local fields of our theory using constraints and gauge conditions (27-30):
$$Z_{\mathrm{local}}(F_1,F_2|P_0,\phi _0,N_0)=\underset{F_1}{\overset{F_2}{}}D(F,P_f)\mathrm{\Delta }_s\overline{\mathrm{\Delta }}_t\mathrm{exp}\left\{i\overline{W}\right\},$$
(58)
where
$$D(F,P_f)=\underset{t,x}{}\left(\underset{i<k}{}\frac{dq^{ik}d\pi _{ik}}{2\pi }\underset{f}{}\frac{dfdp_f}{2\pi }\right)$$
(59)
are functional differentials for the metric fields ($`\pi ,q`$) and the matter fields ($`p_f,f`$),
$$\mathrm{\Delta }_s=\underset{t,x,i}{}\delta (๐ซ_i))\delta (\chi ^j)det\{๐ซ_i,\chi ^j\},$$
(60)
$$\overline{\mathrm{\Delta }}_t=\underset{t,x}{}\delta ((\mu ))\delta (\pi )det\{(\phi _0)\rho ,\pi \},\left(\rho =\frac{d^3xH(\phi _0)}{V_0}\right)$$
(61)
are the F-P determinants, and
$$\overline{W}=\underset{t_1}{\overset{t_2}{}}๐t\left\{\underset{V_0}{}d^3x\left(\underset{F}{}P_F\dot{F}\right)P_0\dot{\phi }_0N_0\left[\frac{P_0^2}{4V_0}+H(\phi _0)\right]+\frac{1}{2}_t(P_0\phi _0)\right\}$$
(62)
is extended action of considered theory.
By analogy with SR considered in Appendix A we define a commutative Green function as an integral over global fields $`(P_0,\phi _0)`$ and the average over reparametrization group parameter $`N_0`$
$$G_+(F_1,\phi _1|F_2,\phi _2)=\underset{\phi _1}{\overset{\phi _2}{}}\underset{t}{}\left(\frac{d\phi _0dP_0d\stackrel{~}{N}_0}{2\pi }\right)Z_{\mathrm{local}}(F_1,F_2|P_0,\phi _0,N_0),$$
(63)
where
$$\stackrel{~}{N}=N/2\pi \delta (0),\delta (0)=๐N_0.$$
(64)
The causal Green function in the world field space ($`F,\phi _0`$) is defined as the sum
$$G_c(F_1,\phi _1|F_2,\phi _2)=G_+(F_1,\phi _1|F_2,\phi _2)\theta (\phi _1\phi _2)+G_+(F_2,\phi _1|F_2,\phi _1)\theta (\phi _2\phi _1).$$
(65)
This function will be considered as generating functional for the unitary $`S`$-matrix elements
$$S[\phi _1,\phi _2]=<\text{o}ut(\phi _2)|T_\phi \mathrm{exp}\left\{i\underset{\phi _1}{\overset{\phi _2}{}}๐\phi (H_I^{})\right\}|(\phi _1)\text{i}n>,$$
(66)
where $`T_\phi `$ is a symbol of ordering with respect to parameter $`\phi _0`$, and $`<\mathrm{out}(\phi _2)|,|(\phi )\mathrm{in}>`$ are states of quantum Univers in the lowest order of the Dirac perturbation theory ($`๐ฉ=1;N^k=0;q^{ij}=\delta _{ij}+h_{ij}^T`$), $`H_I^{}`$ is the interaction Hamiltonian
$$H_I^{}=H^{}H_0^{},H^{}=2\sqrt{V_0H(\phi )},H_0^{}=2\sqrt{V_0H_0(\phi )},$$
(67)
$`H_0`$ is a sum of the Hamiltonians of โfreeโ fields (gravitons, photons, massive vectors, and spinors) where all masses (including the Planck mass) are replaced by the dynamic evolution parameter $`\phi _0`$ . For example for gravitons the โfreeโ hamiltonian takes the form:
$$H_0(\phi _0)=d^3x\left(\frac{6(\pi _{(h)}^T)^2}{\phi _0^2}+\frac{\phi _0^2}{24}(_ih^T)^2\right);(h_{ii}^T=0;_jh_{ji}^T=0).$$
(68)
In order to reproduce Faddeev-Popov integral for general relativity in infinite space-time , one should fix the dynamic evolution parameter at its present-day value $`\phi _0=\mu `$ (38), remove all the zero-mode dynamics $`P_0=\dot{\phi }_0=0,N_0=1`$, and neglect the surface Newton term in the Hamiltonian. We get
$$Z^{FP}(F_1,F_2)=Z_{\mathrm{local}}(F_1,F_2|P_0=0,\phi _{0}^{}{}_{exp}{}^{}=\mu ,N_0=1),$$
(69)
or
$$Z^{FP}(F_1,F_2)=\underset{F_1}{\overset{F_2}{}}D(F,P_f)\mathrm{\Delta }_s\mathrm{\Delta }_t\mathrm{exp}\left\{iW_{fp}\right\},$$
(70)
where
$$W_{fp}=\underset{\mathrm{}}{\overset{+\mathrm{}}{}}๐td^3x\left(\underset{F}{}P_F\dot{F}_{fp}(\mu )\right),_{fp}(\mu )=(\mu )\frac{\mu ^2}{6}_i_jq^{ij},$$
(71)
and
$$\mathrm{\Delta }_t=\underset{t,x}{}\delta ((\mu ))\delta (\pi )det\{(\mu ),\pi \}.$$
(72)
The F-P integral (70) is considered as the generating functional for unitary perturbation theory in terms of S-matrix elements
$$S[\mathrm{}|+\mathrm{}]=<\text{o}ut|T\mathrm{exp}\left\{i\underset{\mathrm{}}{\overset{+\mathrm{}}{}}๐tH_I(\mu )\right\}|\text{i}n>.$$
(73)
Strictly speaking, the approximation (69) is not a correct procedure, as it breaks the reparametrization-invariance. The region of validity of FP integral (70) is discussed in next sections.
## 4. Evolution of โFreeโ Quantum Universe
### 4.1. Dynamic unconstrained system
Possible states of a free quantum universe in $`S`$-matrix (66) are determined by the lowest order of the Dirac perturbation theory given by the well-known system of โfreeโ conformal fields (8), (68) in a finite space-time volume
$$W_0^E=\underset{t_1}{\overset{t_2}{}}๐t\left(\left[d^3x\underset{F}{}P_F\dot{F}\right]P_0\dot{\phi }_0N_0[\frac{P_0^2}{4V}+H_0(\phi _0)]+\frac{1}{2}_0(P_0\phi _0)\right),$$
(74)
where $`H_0`$ is a sum of the Hamiltonians of โfreeโ fields (gravitons (68), photons, massive vectors, and spinors) where all masses (including the Planck mass) are replaced by the dynamic evolution parameter $`\phi _0`$ . The classical equations for the action (74)
$$\frac{dF}{dT}=\frac{H_0}{P_F},\frac{dP_F}{dT}=\frac{H_0}{F},P_0=\pm 2\sqrt{V_0H_0}:=H_0^{}$$
(75)
contain two invariant times: the geometric $`T`$ and the dynamic $`\phi _0^\pm `$ connected by the geometro-dynamic (back-reaction) equation
$$\frac{d\phi _0^\pm }{dT}=\pm \sqrt{\rho _0(\phi _0^\pm )},\left(\rho _0=\frac{H_0}{V_0}\right).$$
(76)
Solving the energy constraint we get the action for dynamic system
$$W_0^E(constraint)=W_0^D=\underset{\phi (t_1)}{\overset{\phi (t_2)}{}}๐\phi \left(\left[d^3x\underset{F}{}P_F_\phi F\right]H_{0\pm }^{}+\frac{1}{2}_\phi (\phi H_{0\pm }^{})\right),$$
(77)
that has two branches for a universe with a positive energy $`(P_0>0)`$, and a universe with a negative energy $`(P_0<0)`$. We interpret the branch with negative energy as an โanti-universeโ which propagates backward $`(\phi <0)`$ with positive energy to provide the stability of a quantum system.
The content of matter in the universe is described by the number of particles $`N_{F,k}`$ and their energy $`\omega _F(\phi _0,k)`$ (which depends on the dynamic evolution parameter $`\phi _0`$ and quantum numbers $`k`$, momenta, spins, etc.). Detected particles are defined as the field variables $`F=f`$
$$f(x)=\underset{k}{}\frac{C_f(\phi _0)\mathrm{exp}(ik_ix_i)}{V_0^{3/2}\sqrt{2\omega _f(\phi _0,k)}}\left(a_f^+(k)+a_f^{}(k)\right)$$
(78)
which diagonalize the operator of the density of matter
$$\rho _0=\underset{f,k}{}\frac{\omega _f(\phi _0,k)}{V_0}\widehat{N}_{f,k},\widehat{N}_f(a)=\frac{1}{2}(a_f^+a_f^{}+a_f^{}a_f^+).$$
(79)
We restrict ourselves to gravitons (f=h) $`C_h(\phi _0)=\phi _0\sqrt{12},\omega _h(\phi _0,k)=\sqrt{k^2}`$ and massive vector particles (f=v) $`C_v(\phi _0)=1,\omega _v(\phi _0,k)=\sqrt{k^2+y^2\phi _0^2}`$, where $`y`$ is the mass in terms of the Planck constant.
### 4.2. Geometric unconstrained system
The equations of motion (76) in terms of $`a^+,a^{}`$ are not diagonal
$$i\frac{d}{dT}\chi :=i\chi _{a_f}^{}=\widehat{H}_{a_f}\chi _{a_f},\chi _{a_f}=\left(\begin{array}{c}a_f^+\\ a_f\end{array}\right);\widehat{H}_{a_f}=\left|\begin{array}{ccc}\omega _{a_f}& ,& i\mathrm{\Delta }_f\\ & & \\ i\mathrm{\Delta }_f& ,& \omega _{a_f}\end{array}\right|,$$
(80)
where nondiagonal terms $`\mathrm{\Delta }_{f=h,v}`$ are proportional to the Hubble parameter (38
$$\mathrm{\Delta }_{f=h}=\frac{\phi _0^{}}{\phi _0},\mathrm{\Delta }_{f=v}=\frac{\omega _v^{}}{2\omega _v},\phi _0^{}=\sqrt{\rho _0}.$$
(81)
The โgeometric systemโ $`(b^+,b)`$ is determined by the transformation to the set of variables which diagonalize equations of motion (75) and determine a set of integrals of motion of equations (75) (as conserved numbers $`\left\{Q\right\}`$).
To obtain integrals of motion and to choose initial conditions for the universe evolution we use the Bogoliubov transformations and define โquasi-particlesโ
$$b^+=\mathrm{cosh}(r)e^{i\theta }a^+i\mathrm{sinh}(r)e^{i\theta }a,b=\mathrm{cosh}(r)e^{i\theta }a+i\mathrm{sinh}(r)e^{i\theta }a^+,$$
(82)
or
$$\chi _b=\left(\begin{array}{c}b^+\\ b\end{array}\right)=\widehat{O}\chi _a,$$
which diagonalize the classical equations expressed in terms of โparticlesโ $`(a^+,a)`$, so that the โnumber of quasiparticlesโ is conserved
$$\frac{d(b^+b)}{dt}=0,b=\mathrm{exp}(i\underset{0}{\overset{T}{}}๐\overline{T}\overline{\omega }_b(\overline{T}))b_0.$$
(83)
Functions $`r`$ and $`\theta `$ in (82), and the quasiparticle energy $`\overline{\omega }_b`$ in (83) are determined by the equation of diagonalization
$$i\frac{d}{dT}\chi _b=[i\widehat{O}{}_{}{}^{1}\frac{d}{dT}\widehat{O}\widehat{O}{}_{}{}^{1}\widehat{H}{}_{a}{}^{}\widehat{O}]\chi _b\left(\begin{array}{cc}\hfill \overline{\omega }_b,& \hfill 0\\ & \\ \hfill 0,& \hfill \overline{\omega }_b\end{array}\right)\chi _b$$
(84)
in the form obtained in
$$\overline{\omega }_{fb}=(\omega _f\theta _f^{})\mathrm{cosh}(2r_f)(\mathrm{\Delta }_f\mathrm{cos}2\theta _f)\mathrm{sinh}(2r_f),$$
(85)
$$0=(\omega _f\theta _f^{})\mathrm{sinh}(2r_f)(\mathrm{\Delta }_f\mathrm{cos}2\theta _f)\mathrm{cosh}(2r_f),r_f^{}=\mathrm{\Delta }_f\mathrm{sin}2\theta _f.$$
Equations (81)โ (85) are closed by the definition of โobservable particlesโ in terms of quasiparticles
$$\rho (\phi )=\frac{H_0}{V}=\frac{\underset{f}{}\omega _f(\phi )\{a_f^+a_f\}}{V_0};\{a^+a\}=\{b_0^+b_0\}\mathrm{cosh}2r\frac{i}{2}(b^{+2}b^2)\mathrm{sinh}2r$$
(86)
with
$$\overline{\omega }_{fb}=\sqrt{(\omega _f\theta _f^{})^2+(r_f^{})^2\mathrm{\Delta }_f^2},\theta _f^{}=\frac{1}{2}\left(\frac{r_f^{}}{\mathrm{\Delta }_f}\right)^{}[1\frac{(r_f^{})^2}{\mathrm{\Delta }_f^2}]^{1/2},\mathrm{cosh}(2r_f)=\frac{\omega _f\theta _f^{}}{\overline{\omega }_{fb}}.$$
(87)
The constrained system in terms of geometric variables is described by the action
$$\stackrel{~}{W}^G=๐t\left\{\underset{f}{}\frac{i}{2}\left(b_tb^+b^+_tb\right)_f\stackrel{~}{\mathrm{\Pi }}_0\dot{Q}_0N_0\left[\stackrel{~}{\mathrm{\Pi }}_0+\underset{f}{}\omega _b^f(Q_0)N_f(b)\right]\right\},$$
(88)
where the new dynamic evolution parameter $`Q_0`$ coincides with geometric time $`T`$ on the equations of motion
$$\frac{\delta \stackrel{~}{W}^E}{\delta \stackrel{~}{\mathrm{\Pi }}_0}=0\frac{dQ_0}{dt}dQ_0=dT.$$
(89)
Reduction of this system leads to the weak version of Geometric Unconstrained System (49)
$$\stackrel{~}{W}^{GUS}=๐T\left\{\underset{f}{}\frac{i}{2}\left(b_Tb^+b^+_Tb\right)_f\underset{f}{}\omega _b^f(T)N_f(b)\right\}.$$
(90)
We choose the initial data appropriate for the dynamics described by GUS (90).
### 4.3. Quantization
The initial data $`b_0,b_0^+`$ of quasiparticle variables (83) form the set of quantum numbers in quantum theory.
Let us suppose that we manage to solve equations (83)โ (87) with respect to the geometric time $`T`$ in terms of conserved numbers $`b_0^+,b_0`$. This means that the wave function of a quantum universe can be represented in the form of a series over the conserved quantum numbers $`Q=n_{f,k}=<Q|b_f^+b_f|Q>`$ of the Bogoliubov states
$$\mathrm{\Phi }_Q(T)=\underset{f,n_f}{}\mathrm{exp}\left\{i\underset{0}{\overset{T}{}}๐Tn_f\overline{\omega }_b(T)\right\}\frac{(b_f^+)^{n_f}}{\sqrt{n_f!}}|0>_b.$$
(91)
In this geometric system, we have an arrow of the geometric time $`T`$ for the universe
$$T_+(\phi _2,\phi _1)=\underset{\phi _1}{\overset{\phi _2}{}}๐\phi \rho (\phi )^{1/2}>0,\phi _2>\phi _1,$$
(92)
and for anti-universe
$$T_{}(\phi _2,\phi _1)=\underset{\phi _1}{\overset{\phi _2}{}}๐\phi \rho (\phi )^{1/2}=\underset{\phi _2}{\overset{\phi _1}{}}๐\phi \rho (\phi )^{1/2}>0,\phi _1>\phi _2.$$
(93)
The dynamic system (77) of particle variables $`a^+,a`$ is connected with the geometric one by the Bogoliubov transformations. Using these transformations we can find wave functions of a universe, for $`\phi _2>\phi _1`$ and an anti-Universe, for $`\phi _1>\phi _2`$
$$\mathrm{\Psi }_Q(T)=A_Q^+\mathrm{\Phi }_Q(T_+(\phi _2,\phi _1))\theta (\phi _2\phi _1)+A_Q^{}\mathrm{\Phi }_Q^{}(T_{}(\phi _2,\phi _1))\theta (\phi _1\phi _2),$$
(94)
where the first term and the second one are positive ($`P_0>0`$) and negative ($`P_0<0`$) frequency parts of the solutions with the spectrum of quasiparticles $`\overline{\omega }_b`$, $`A_Q^+`$ is the operator of creation of a universe with a positive โfrequencyโ (which propagates in the positive direction of the dynamic evolution parameter) and $`A_Q^{}`$ is the operator of annihilation of a universe (or creation of an anti-universe) with a negative โfrequencyโ (which propagates in the negative direction of the dynamic evolution parameter).
We can see that the creation of the universe in the field world space and the creation of dynamic particles by the geometric vacuum $`(b^+|0>=0)`$ are two different effects.
The second effect disappears if we neglect gravitons and massive fields. In this case, $`d\rho /d\phi =0`$, and one can represent a wave function of Universe in the form of the spectral series over eigenvalues $`\rho _Q`$ of the density $`\rho `$
$$\mathrm{\Psi }(f|\phi _2,\phi _1)=$$
(95)
$$\underset{Q}{}\left[\frac{A_Q^+}{\sqrt{2\rho _Q}}\mathrm{exp}\left\{i(\phi _2\phi _1)\frac{\overline{\omega }_fn_f}{\sqrt{\rho _Q}}\right\}<f|Q>+\frac{A_Q^{}}{\sqrt{2\rho _Q}}\mathrm{exp}\left\{i(\phi _2\phi _1)\frac{\overline{\omega }_fn_f}{\sqrt{\rho _Q}}\right\}<f|Q>^{}\right],$$
where $`<f|Q>`$ is a product of normalizable Hermite polynomials.
### 4.4. Evolution of quantum universe
The equations of diagonalization (84) for the Bogoliubov coefficients (82) and the quasiparticle energy $`\overline{\omega }_b`$ (85) play the role of the equations of state of the field matter in the universe. We can show that the choice of initial conditions for the โBig Bangโ in the form of the Bogoliubov (squeezed) vacuum $`b|0>_b=0`$ reproduces all stages of the evolution of the Friedmann-Robertson-Walker universe in their conformal versions: anisotropic, inflation, radiation, and dust.
The squeezed vacuum (i.e., the vacuum of quasiparticles) is the state of โnothingโ. For small $`\phi `$ and a large Hubble parameter, at the beginning of the universe, the state of vacuum of quasiparticles leads to the density of matter
$${}_{b}{}^{}<\rho (a^+,a)>_b=\rho _0\frac{1}{2}(\frac{\phi ^2(0)}{\phi ^2(T)}+\frac{\phi ^2(T)}{\phi ^2(0)}),\theta =\frac{\pi }{4}$$
(96)
where $`\phi (0)`$ is the initial value, and $`\rho _0`$ is the density of the Casimir energy of vacuum of โquasiparticlesโ. The first term corresponds to the conformal version of the rigid state equation (in accordance with the classification of the standard cosmology) which describes the Kasner anisotropic stage $`T_\pm (\phi )\pm \phi ^2`$ (considered on the quantum level by Misner ). The second term of the squeezed vacuum density (96) (for an admissible positive branch) leads to the stage with inflation of the dynamic evolution parameter $`\phi `$ with respect to the geometric time $`T`$
$$\phi (T)_{(+)}\phi (0)\mathrm{exp}[T\sqrt{2\rho _0}/\phi (0)].$$
It is the stage of intensive creation of โmeasurable particlesโ. After the inflation, the Hubble parameter goes to zero, and gravitons convert into photon-like oscillator excitations with the conserved number of particles.
At the present-day stage, the Bogoliubov quasiparticles coincide with particles, so that the measurable density of energy of matter in the universe is a sum of relativistic energies of all particles
$$\rho _0(\phi )=\frac{E}{V_0}=\underset{n_f}{}\frac{n_f}{V_0}\sqrt{k_{fi}^2+y_f^2\phi ^2(T)},$$
(97)
where $`y_f`$ is the mass of a particle in units of the Planck mass. The case of massless particles ($`y=0,\rho _0(\phi )=constant`$) correspond to the conformal version of radiation stage of the standard FRW-cosmology. And the massive particles at rest ($`k=0,\rho _0(\phi )=\rho _{\mathrm{barions}}\phi /\mu )`$ corresponds to the conformal version of the dust universe of the standard cosmology with the Hubble law
$$\phi ^{}=\pm \sqrt{\rho _0}\phi _\pm (T)=\left(\frac{\rho _{\mathrm{barions}}}{4\mu }\right)T^2;q=\frac{\phi ^{\prime \prime }\phi }{\phi _{}^{}{}_{}{}^{2}}=\frac{1}{2}.$$
(98)
The dynamic evolution parameter is expressed through the geometric time of a quantum asymptotic state of the universe $`|out>`$ and conserved quantum numbers of this state: energy $`E_{out}`$ and density $`\rho _0=E_{out}/V_0`$.
It is well-known that $`E_{out}`$ is a tremendous energy $`(10^{79}GeV)`$ in comparison with possible real and virtual deviations of the free Hamiltonian in the laboratory processes:
$$\overline{H}_0=E_{out}+\delta H_0,<\text{o}ut|\delta H_0|\text{i}n><<E_{out}.$$
(99)
We have seen that the dependence of the scale factor $`\phi _0`$ on the geometric time $`T`$ (or the โrelationโ of two classical unconstrained systems: dynamic and geometric) describes the โBig Bangโ and evolution of a universe.
Therefore, from the point of view of unconstrained system โBig Bangโ is effect of evolution of the geometric interval with respect to the dynamic evolution parameter which goes beyond the scope of Hamiltonian description of a single classical unconstrained system.
Reparametrization-invariant dynamic of General Relativity is covered by Geometric and Dynamic Unconstrained Systems connected by the Levi-Civita transformation of fields of matter into the vacuum fields of initial data with respect to geometric time (see Fig. 3).
## 5. QFT limit of Quantum Gravity
The simplest way to determine the QFT limit of Quantum Gravity and to find the region of validity of the FP-integral(70) is to use the quantum field version of the reparametrization-invariant integral(63) in the form of S-matrix elements (see (66), (67)). We consider the infinite volume limit of the S-matrix element (67) in terms of the geometric time $`T`$ for the present-day stage $`T=T_0,\phi (T_0)=\mu `$, and $`T(\phi _1)=T_0\mathrm{\Delta }T,T(\phi _2)=T_0+\mathrm{\Delta }T=T_{out}`$. One can express this matrix element in terms of the time measured by an observer of an out-state with a tremendous number of particles in the Universe using equation (98) $`d\phi =dT_{out}\sqrt{\rho _{out}}`$ and approximation (99) to neglect โback-reactionโ. In the infinite volume limit, we get from (67)
$$d\phi _0[H_I^{}]=2d\phi _0\left(\sqrt{V_0(H_0+H_I)}\sqrt{V_0H_0}\right)=dT_{out}[\widehat{F}\overline{H}_I+O(1/E_{out})]$$
(100)
where $`H_I`$ is the interaction Hamiltonian in GR, and
$$\widehat{F}=\sqrt{\frac{E_{out}}{H_0}}=\sqrt{\frac{E_{out}}{E_{out}+\delta H_0}}$$
(101)
is a multiplier which plays the role of a form factor for physical processes observed in the โlaboratoryโ conditions when the cosmic energy $`E_{out}`$ is much greater than the deviation of the free energy
$$\delta H_0=H_0E_{out};$$
(102)
due to creation and annihilation of real and virtual particles in the laboratory experiments.
The measurable time of the laboratory experiments $`T_2T_1`$ is much smaller than the age of the universe $`T_0`$, but it is much greater than the reverse โlaboratoryโ energy $`\delta `$, so that the limit
$$\underset{T(\phi _1)}{\overset{T(\phi _2)}{}}๐T_{\mathrm{out}}\underset{\mathrm{}}{\overset{+\mathrm{}}{}}๐T_{\mathrm{out}}$$
is valid. If we neglect the form factor (101) that removes a set of ultraviolet divergences, we get the matrix element (73) that corresponds to the standard FP functional integral (70) and $`S`$-matrix element (73) where the coordinate time is replaced by the geometric (conformal) time $`tT_{\mathrm{out}}`$:
$$S[\mathrm{}|+\mathrm{}]=<\text{o}ut|T\mathrm{exp}\left\{i\underset{\mathrm{}}{\overset{+\mathrm{}}{}}๐T_{\mathrm{out}}\widehat{F}H_I(\mu )\right\}|\text{i}n>\left(\widehat{F}=1\right).$$
(103)
Thus, the standard FP-integral (i.e., the Hamiltonian description of the evolution of fields with respect to the geometric time) appears as the nonrelativistic approximation of tremendous mass of a universe and its very large life-time (see Fig. 4). Now, it is evident that the conventional FP unitary S-matrix are not valid for the description of the early universe given in the finite spatial volume and the finite positive interval of geometrical time when the early universe only begins to create matter.
On the other hand, we revealed that standard quantum field theory (103) is expressed in terms of the conformal-invariant Lichnerowicz variables and coordinates including the conformal time ($`T_{out}`$) as the time of evolution of these variables. The reparametrization-invariant description of the โBig Bangโ evolution distinguishes conformal variables (8) and coordinates. The conformal invariance of the variables can testify to the conformal invariance of the initial theory of gravity. Such the theory can be a scalar version of the Weyl conformal invariant theory (that dynamically coincides with the Einstein General Relativity) where these conformal variables, coordinates, geometric time $`T`$, and the conformal Hubble parameter $`_{hub}^c=\phi ^{}/\phi `$ can be considered as measurable quantities. In particular, to get โacceleratingโ universe with $`q>0`$ in the dust stage (see (98)), is enough to count that we measure the relative Weyl time ($`T_{\mathrm{out}}`$) of quantum field theory (103).
## 6. Conclusions
The main result of the paper is the reparametrization-invariant generating functional for the unitary and causal perturbation theory in general relativity in a finite space-time. We show that the classical cosmology of a universe and the Faddeev-Popov-DeWitt functional correspond to different orders of decomposition of this functional over the inverse โmassโ of a universe.
This result is based on the assertion that the measurable time in any reparametrization-invariant system is not the coordinate time, but the time-like dynamic variable of an extended phase space of this system (of the type of the conformal scale factor). Accordingly, the measurable Hamiltonian is a solution of the energy constraint with respect to the conjugate momentum of this dynamic evolution parameter . This definition of the dynamic evolution parameter and Hamiltonian supposes the reduction of an action for constructing an โequivalentโ unconstrained system.
The second assertion is that such an unconstrained dynamic system cannot cover the physical content of a relativistic reparametrization-invariant system. This content can be covered by two โequivalentโ unconstrained systems โ the mentioned above dynamic system and the geometric system constructed by the Levi-Civita type canonical transformation so that a new dynamic evolution parameter coincides with the geometric time.
The โgeometricโ variables are the Bogoliubov quasiparticles which diagonalize equations of motion and give cosmological initial conditions. The coefficients of the Bogoliubov transformations, in the conventional QFT perturbation theory, determine the evolution density of matter. For the vacuum initial data this evolution reproduces all stages of the standard FRW evolution of the universe in their conformal versions.
Consistent QFT limits of the generating functional in classical gravitation and cosmology and the conventional quantum field theory (in the form of the Faddeev-Popov generating functional for a infinite space-time) can be fulfilled for a tremendous value of the universe mass and the universe life-time (see Fig. 4). The quantum field theory limit distinguish the conformal treatment of general relativity developed in .
Acknowledgments
We are happy to acknowledge interesting and critical discussions with Profs. B. M. Barbashov, A. Borowiec, A. M. Chervyakov, A.V. Efremov, G.A. Gogilidze, V.G. Kadyshevsky, A.M. Khvedelidze, H. Kleinert, E.A. Kuraev, D. Mladenov, V.V. Papoyan, V. I. Smirichinski, J. ลniatycki, I. V. Tyutin, Yu. S. Vladimirov. The work was supported by the Committee for Scientific Research grant no. 603/P03/96 and by the Infeld-Bogoliubov program.
APPENDIX A. Special Relativity
To answer the question: Why is the reparametrization-invariant Hamiltonian reduction needed?, let us consider relativistic mechanics in the Hamiltonian form
$$W[P,X|N|\tau _1,\tau _2]=\underset{\tau _1}{\overset{\tau _2}{}}๐\tau [P_\mu \dot{X}^\mu \frac{N}{2m}(P_\mu ^2+m^2)].$$
(A.1)
This action is invariant with respect to reparametrizations of the coordinate evolution parameter
$$\tau \tau ^{}=\tau ^{}(\tau ),N^{}d\tau ^{}=Nd\tau $$
(A.2)
given in the one-dimensional space with the invariant interval
$$dT:=Nd\tau ,T=\underset{0}{\overset{\tau }{}}๐\overline{\tau }N(\overline{\tau }).$$
(A.3)
We called this invariant interval the geometric time whereas the dynamic variable $`X_0`$ (with a negative contribution in the constraint) we called dynamic evolution parameter.
In terms of the geometric time (A.3) the classical equations of the generalized Hamiltonian system (A.1) takes the form
$$\frac{dX_\mu }{dT}=\frac{P_\mu }{m},\frac{dP_\mu }{dT}=0,P_\mu ^2m^2=0.$$
(A.4)
The problem is to obtain the equivalent unconstrained theories directly in terms of the invariant times $`X_0`$ or $`T`$ with the invariant Hamiltonians of evolution with respect to these times. The solution of this problem is called the dynamic (for $`X_0`$), or geometric (for $`T`$) reparametrization-invariant Hamiltonian reductions.
The dynamic reduction of the extended system (A.1) means the substitution, into it, of the explicit resolving of the energy constraint $`(P_\mu ^2+m^2)=0`$ with respect to the momentum $`P_0`$ with a negative contribution
$$\frac{\delta W}{\delta N}=0P_0=\pm \sqrt{m^2+P_i^2}.$$
(A.5)
In accordance with the two signs of the solution (A.5), after the substitution of (A.5) into (A.1), we have two branches of the dynamic unconstrained system
$$W(\text{constraint})=W_\pm ^D[P_i,X_i|X_0(1),X_0(2)]=\underset{X_0(\tau _1)=X_0(1)}{\overset{X_0(\tau _2)=X_0(2)}{}}๐X_0\left[P_i\frac{dX_i}{dX_0}\sqrt{m^2+P_i^2}\right].$$
(A.6)
The role of the time of evolution, in this action, is played by the variable $`X_0`$ which abandons the Dirac sector of โobservablesโ $`P_i,X_i`$, but not the sector of โmeasurableโ quantities. At the same time, its conjugate momentum $`P_0`$ converts into the corresponding Hamiltonian of evolution, values of which are the energy of a particle.
This invariant reduction of the action gives the โequivalentโ unconstrained system together with definition of the invariant evolution parameter ($`X_0`$) corresponding to a non-zero Hamiltonian $`P_0`$.
Thus, we need the reparametrization-invariant Hamiltonian reduction to determine the invariant evolution parameter and its invariant Hamiltonian for reparametrization-invariant systems.
In quantum relativistic theory, we get two Schrรถdinger equations
$$i\frac{d}{dX_0}\mathrm{\Psi }_{(\pm )}(X|P)=\pm \sqrt{m^2+P_i^2}\mathrm{\Psi }_{(\pm )}(X|P),$$
(A.7)
with positive and negative values of $`P_0`$ and normalized wave functions
$$\mathrm{\Psi }_\pm (X|P)=\frac{A_P^\pm \theta (\pm P_0)}{(2\pi )^{3/2}\sqrt{2P_0}}\mathrm{exp}(iP_\mu X^\mu ),\left([A_P^{},A_P^{}^+]=\delta ^3(P_iP_i^{})\right).$$
(A.8)
The coefficient $`A_P^+`$, in the secondary quantization, is treated as the operator of creation of a particle with positive energy; and the coefficient $`A_P^{}`$, as the operator of annihilation of a particle also with positive energy. The physical states are formed by action of these operators on the vacuum $`<0|,|0>`$ in the form of out-state ( $`|P>=A_P^+|0>`$ ) with positive frequencies and in-state ( $`<P|=<0|A_P^{}`$ ) with negative frequencies. This treatment means that positive frequencies propagate forward ($`X_{0}^{}{}_{2}{}^{}>X_{0}^{}{}_{1}{}^{}`$); and negative frequencies, backward ($`X_{0}^{}{}_{1}{}^{}>X_{0}^{}{}_{2}{}^{}`$), so that the negative values of energy are excluded from the spectrum to provide the stability of the quantum system in QFT . For this causal convention the geometric time (A.3) is always positive in accordance with the equations of motion (A.4)
$$\left(\frac{dT}{dX_0}\right)_\pm =\pm \frac{m}{\sqrt{P_i^2+m^2}}T(X_{0}^{}{}_{2}{}^{},X_{0}^{}{}_{1}{}^{})=\pm \frac{m}{\sqrt{P_i^2+m^2}}(X_{0}^{}{}_{2}{}^{}X_{0}^{}{}_{1}{}^{})0$$
(A.9)
In other words, instead of changing the sign of energy, we change that of the dynamic evolution parameter, which leads to the arrow of the geometric time (A.9) and to the causal Green function
$$G^c(X)=G_+(X)\theta (X_0)+G_{}(X)\theta (X_0)=i\frac{d^4P}{(2\pi )^4}\mathrm{exp}(iPX)\frac{1}{P^2m^2iฯต},$$
(A.10)
where $`G_+(X)=G_{}(X)`$ is the โcommutativeโ Green function
$$G_+(X)=\frac{d^4P}{(2\pi )^3}\mathrm{exp}(iPX)\delta (P^2m^2)\theta (P_0)=\frac{1}{2\pi }d^3Pd^3P^{}<0|\mathrm{\Psi }_{}(X|P)\mathrm{\Psi }_+(0|P^{})|0>.$$
(A.11)
To obtain the reparametrization-invariant form of the functional integral adequate to the considered gauge-less reduction (A.6) and the causal Green function (A.10), we use the version of composition law for the commutative Green function with the integration over the whole measurable sector $`X_{1\mu }`$
$$G_+(XX_0)=G_+(XX_1)\overline{G}_+(X_1X_0)๐X_1(\overline{G}_+=\frac{G_+}{2\pi \delta (0)}),$$
(A.12)
where $`\delta (0)=๐N`$ is the infinite volume of the group of reparametrizations of the coordinate $`\tau `$. Using the composition law $`n`$-times, we got the multiple integral
$$G_+(XX_0)=G_+(XX_1)\underset{k=1}{\overset{n}{}}\overline{G}_+(X_kX_{k+1})dX_k,(X_{n+1}=X_0).$$
(A.13)
The continual limit of the multiple integral with the integral representation for $`\delta `$-function
$$\delta (P^2m^2)=\frac{1}{2\pi }๐N\mathrm{exp}[iN(P^2m^2)]$$
can be defined as the path integral in the form of the average over the group of reparametrizations
$$G_+(X)=\underset{X(\tau _1)=0}{\overset{X(\tau _2)=X}{}}\frac{dN(\tau _2)d^4P(\tau _2)}{(2\pi )^3}\underset{\tau _1\tau <\tau _2}{}\left\{d\overline{N}(\tau )\underset{\mu }{}\left(\frac{dP_\mu (\tau )dX_\mu (\tau )}{2\pi }\right)\right\}\mathrm{exp}(iW[P,X|N|\tau _1,\tau _2]),$$
(A.14)
where $`\overline{N}=N/2\pi \delta (0)`$, and $`W`$ is the initial extended action (A.1).
This functional integral has the form of the average over the group of reparametrization of the integral over the sector of โmeasurableโ variables $`P_\mu ,X_\mu `$.
The Hamiltonian unconstrained system in terms of the geometric time $`T`$ can be obtained by the canonical Levi-Civita - type transformation
$$(P_\mu ,X_\mu )(\mathrm{\Pi }_\mu ,Q_\mu )$$
(A.15)
to the variables ($`\mathrm{\Pi }_\mu ,Q_\mu `$) for which one of equations identifies $`Q_0`$ with the geometric time $`T`$. This transformation converts the constraint into a new momentum
$$\mathrm{\Pi }_0=\frac{1}{2m}[P_0^2P_i^2],\mathrm{\Pi }_i=P_i,Q_0=X_0\frac{m}{P_0},Q_i=X_iX_0\frac{P_i}{P_0}$$
(A.16)
and has the inverted form
$$P_0=\pm \sqrt{2m\mathrm{\Pi }_0+\mathrm{\Pi }_i^2},P_i=\mathrm{\Pi }_i,X_0=\pm Q_0\frac{\sqrt{2m\mathrm{\Pi }_0+\mathrm{\Pi }_i^2}}{m},X_i=Q_i+Q_0\frac{\mathrm{\Pi }_i}{m}.$$
(A.17)
After transformation (A.16) the action (A.1) takes the form
$$W=\underset{\tau _1}{\overset{\tau _2}{}}๐\tau \left[\mathrm{\Pi }_\mu \dot{Q}^\mu N(\mathrm{\Pi }_0+\frac{m}{2})\frac{d}{d\tau }S^{lc}\right],S^{lc}=(Q_0\mathrm{\Pi }_0).$$
(A.18)
The invariant reduction is the resolving of the constraint $`\mathrm{\Pi }_0=m/2`$ which determines a new Hamiltonian of evolution with respect to the new dynamic evolution parameter $`Q_0`$, whereas the equation of motion for this momentum $`\mathrm{\Pi }_0`$ identifies the dynamic evolution parameter $`Q_0`$ with the geometric time $`T`$ ($`Q_0=T`$). The substitution of these solutions into the action (A.18) leads to the reduced action of a geometric unconstrained system
$$W(\text{constraint})=W^G[\mathrm{\Pi },Q_i|T_1,T_2]=\underset{T_1}{\overset{T_2}{}}๐T\left(\mathrm{\Pi }_i\frac{dQ_i}{dT}\frac{m}{2}\frac{d}{dT}(S^{lc})\right)(S^{lc}=Q_0\frac{m}{2}),$$
(A.19)
where variables $`\mathrm{\Pi }_i,Q_i`$ are cyclic ones and have the meaning of initial conditions in the comoving frame
$$\frac{\delta W}{\delta \mathrm{\Pi }_i}=\frac{dQ_i}{d\tau }=0Q_i=Q_i^{(0)},\frac{\delta W}{\delta Q_i}=\frac{d\mathrm{\Pi }_i}{d\tau }=0\mathrm{\Pi }_i=\mathrm{\Pi }_i^{(0)}.$$
(A.20)
The substitution of all geometric solutions
$$Q_0=T,\mathrm{\Pi }_0=\frac{m}{2},\mathrm{\Pi }_i=\mathrm{\Pi }_i^{(0)}=P_i,Q_i=Q_i^{(0)}$$
(A.21)
into the inverted Levi-Civita transformation (A.17) leads to the conventional relativistic solution for the dynamical system
$$P_0=\pm \sqrt{m^2+P_i^2},P_i=\mathrm{\Pi }_i^{(0)},X_0(T)=T\frac{P_0}{m},X_i(T)=X_i^{(0)}+T\frac{P_i}{m}.$$
(A.22)
The Schrรถdinger equation for the wave function
$$\frac{d}{idT}\mathrm{\Psi }^{lc}(T,Q_i|\mathrm{\Pi }_i)=\frac{m}{2}\mathrm{\Psi }^{lc}(T,Q_i|\mathrm{\Pi }_i),\mathrm{\Psi }^{lc}(T,Q_i|\mathrm{\Pi }_i)=\mathrm{exp}(iT\frac{m}{2})\mathrm{exp}(i\mathrm{\Pi }_i^{(0)}Q_i)$$
(A.23)
contains only one eigenvalue $`m/2`$ degenerated with respect to the cyclic momentum $`\mathrm{\Pi }_i`$. We see that there are differences between the dynamic and geometric descriptions. The dynamic evolution parameter is given in the whole region $`\mathrm{}<X_0<+\mathrm{}`$, whereas the geometric one is only positive $`0<T<+\mathrm{}`$, as it follows from the properties of the causal Green function (A.10) after the Levi-Civita transformation (A.16)
$$G^c(Q_\mu )=\underset{\mathrm{}}{\overset{+\mathrm{}}{}}d^4\mathrm{\Pi }_\mu \frac{\mathrm{exp}(iQ^\mu \mathrm{\Pi }_\mu )}{2m(\mathrm{\Pi }_0m/2iฯต/2m)}=\frac{\delta ^3(Q)}{2m}\theta (T),T=Q_0.$$
Two solutions of the constraint (a particle and antiparticle) in the dynamic system correspond to a single solution in the geometric system.
Thus, the reparametrization-invariant content of the equations of motion of a relativistic particle in terms of the geometric time is covered by two โequivalentโ unconstrained systems: the dynamic and geometric. In both the systems, the invariant times are not the coordinate evolution parameter, but variables with the negative contribution into the energy constraint. The Hamiltonian description of a relativistic particle in terms of the geometric time can be achieved by the Levi-Civita-type canonical transformation, so that the energy constraint converts into a new momentum. Whereas, the dynamic unconstrained system is suit for the secondary quantization and the derivation of the causal Green function that determine the arrow of the geometric time.
FIGURE CAPTIONS
Fig. 1. The tree ot modern theoretical physics grew from two different roots (โparticleโ and โfieldโ) which gave the VARIATIONAL method and SYMMETRY principles for formulating modern physical theories as constraned systems. To obtain unambigious physical results, one should construct Equivalent Unconstrained Systems compatible with the symplest variational method. It is just the problem discussed in the present paper.
Fig. 2. An equivalent unconstrained system $`W^{}(p^{},q^{})`$ can be obtained in the case when the operations of varying and constraining commute with each other. The next problem is to establish the range of validity of the standard Faddeev-Popov (FP) integral.
Fig. 3. Reparametrization-invariant dynamics of General Relativity is covered by the Dynamic Unconstrained Systems (DUS) and the Geometric Unconstrained Systems (GUS) connected by the Levi-Civita transformation of fields of MATTER into the vacuum fields of initial data with respect to geometric TIME.
Fig. 4. Reparametrization-invariant dynamics of General Relativity. The Big Bang of Quantum Universe from point of view of Geometric and Dynamic Unconstrained Systems connected by Levi-Civita canonical transformations.
|
warning/0006/hep-th0006214.html
|
ar5iv
|
text
|
# Dynamics and causality constraints
## I Introduction
Causality implementation in field theory is naturally connected to the very concept of field propagation. Old and well known problems appear with a field in a close neighbourhood to its sources. Then a careful analysis is required as the kinematical constraint of a causal propagation is mixed with the dynamics of the field-source interaction. In particular, for a point-like source, there are problems with infinities and other signs of inconsistency. Thus there is a generalized belief that these infinities are consequences of the source point-size dimension and, consequently, that a point-particle cannot be regarded as a viable model for a charged elementary physical object. This, as shown in , does not correspond to reality. The infinities associated to a point-charge self-field are consequences of the way causality has being implemented with the use of lightcones, whose vertex is a singular point; the field infinity just reflects this singularity. This work returns to the ideas raised in further discussing its physical and geometrical meanings. Although it is being based on the case of a point electric charge, its conclusions are of a wider generality, being valid not only for classical field theory but for any theory of fields defined with support on a conic hypersurface (Quantum Field Theory, Quantum Mechanics, General Relativity, Statistical Mechanics, etc), i.e. any theoretic frame work with causality implementation. It shows that the plain Maxwellโs theory, in a short-distance limit, reveals unequivocal and previously unsuspected signs of a quantum nature (the existence of photons) through the indication of a discreteness on the electromagnetic interaction, hidden behind the classical continuous formalism. It hints a proposal of a new approach (fully developed elsewhere where fields and sources are symmetrically treated as discrete objects from which the standard continuous fields are retreated as spacetime effective averages.
This paper is organized in the following way. The geometric vision of causality is discussed in Section II with the introduction of the new concept of extended causality in contraposition to the usual local causality and of their connection to wave-particle duality. Their relevance to point-charge electrodynamics and their inherent conflicts are discussed in Section III. The implications of extended causality on field-source dynamics is exposed in Sections IV and V. The paper ends with some final comments and the conclusions in Section VI.
## II Causality and spacetime geometry
The notation used is of omitting the spacetime indices when this causes no ambiguity. For example, $``$ for $`_\mu `$, and $`A(x,\tau )`$ for a vector field $`A^\mu (x,\tau )`$; $`x`$ stands for both, the event parameterized by $`x^\mu =(t,\stackrel{}{x})`$ and for the coordinate $`x^\mu `$ itself.
The propagation of a massless field on a flat spacetime of metric $`\eta _{\mu \nu }=diag(1,1,1,1),`$ is restricted by
$$\mathrm{\Delta }x^2=0,$$
(1)
which defines a local double (past and future) lightcone: $`\mathrm{\Delta }t=\pm |\mathrm{\Delta }\stackrel{}{x}|.`$ This is also a mathematical expression of local causality in the sense that it is a restriction for the massless field to remain on this lightcone. It is a particular case of the more generic expression
$$\mathrm{\Delta }\tau ^2=\mathrm{\Delta }x^2,$$
(2)
which is, besides, the definition of the proper time $`\tau `$ associated to the propagation of a free physical object across $`\mathrm{\Delta }x`$. As $`\tau `$ is a real valued parameter, the eq. (2) just expresses that $`\mathrm{\Delta }x`$ cannot be space like. Geometrically it is also the definition of a three-dimensional double cone, of which the lightcone and the time axis are just the two extreme limiting cases. $`\mathrm{\Delta }x`$ is the four-vector separation between a generic event $`x`$ and the hypercone vertex. This conic hypersurface is the support for the definition of a propagating field. The hypercone aperture-angle $`\theta ,\mathrm{\hspace{0.33em}\hspace{0.33em}0}\theta \pi /4,`$ is given by $`\mathrm{tan}\theta =\frac{|\mathrm{\Delta }\stackrel{}{x}|}{|\mathrm{\Delta }t|},c=1,`$ or $`\mathrm{\Delta }\tau ^2=(\mathrm{\Delta }t)^2(1\mathrm{tan}^2\theta ).`$ A change of the supporting hypercone corresponds to a change of speed of propagation and is an indication of interaction.
On requirements of continuity one must consider the constraint (2) on a neighbourhood of $`x`$: $`(\mathrm{\Delta }\tau +d\tau )^2=(\mathrm{\Delta }x+dx)^2`$ or, after using eq. (2), $`\mathrm{\Delta }\tau d\tau +\mathrm{\Delta }x.dx=0`$, which may be written as
$$d\tau +f.dx=0,$$
(3)
where $`f`$ is a four-vector tangent to the hypercone (2). For $`\mathrm{\Delta }\tau 0`$ it is just
$$f^\mu :=\frac{\mathrm{\Delta }x^\mu }{\mathrm{\Delta }\tau }|_{\mathrm{\Delta }\tau ^2+\mathrm{\Delta }x^2=0}$$
(4)
For $`\mathrm{\Delta }\tau =0`$ the hypercone (2) reduces to the lightcone (1) and $`f`$ to its tangent four-vector; $`f`$ and $`\mathrm{\Delta }x`$ are both lightlike. It is important to observe that $`f`$ is well defined for any $`\mathrm{\Delta }\tau `$, including $`\mathrm{\Delta }\tau =0`$, as long as $`\mathrm{\Delta }x0.`$ A tangent is not defined at the hypercone vertex. This is a crux point, neglected in the existing literature which leads to the old and well known vexing problems of consistency in classical elctrodynamics . Geometrically eq. (3) defines a hyperplane tangent to the hypercone (2). The simultaneous imposition of eqs. (2) and (3) on the propagation of a free point object produces a much more stringent constraint than local causality as the object is restricted to remain on the intersection of the hypercone (2) with its tangent hyperplane (3), that is, on the hypercone generator tangent to $`f`$, or the $`f`$-generator, for short. This corresponds to an extended concept of causality which will be referred as extended causality.
Local and extended causality correspond to two distinct and complementary (like geometric and wave optics) description of a same physical system. They correspond to different perceptions of the spacetime available to the free evolution of a physical system from a given initial condition, respectively as foliations of hypercones and as congruences of straight lines, the hypercone generators. So, whereas the first one is appropriated for a description in terms of continuous and extended objects like a fluid, a field, a wave, the second one implies on a perception of them as sets of points, describing individually each point. Failure on recognizing this leads to problems of consistency in field theory at short distance.
## III Point-charge electrodynamics
Consider, for example, $`z(\tau )`$, the worldline of a classical point electron parameterized by its proper time $`\tau ;`$ each event on this worldline belongs to the (instantaneous) hypercone
$$\mathrm{\Delta }\tau ^2+\mathrm{\Delta }z^2=0$$
and the four-vector $`u=\frac{dz}{d\tau }`$ is tangent to the worldline (and to the hypercone). It satisfies
$$d\tau +u.dz=0,$$
which corresponds to eq. (3). A free electron remains on the $`u`$-generator of its hypercone; an accelerated electron is on a $`u`$-generator of its instantaneous hypercone. So, in a way, classical electrodynamics already uses extended causality for specifying the state of the classical electron, and this is consistent with an electron modeled as a point particle. This work discusses how extended causality is also used in the definition of its electromagnetic field and the conflicting problems brought with this, pointing the way to a new consistent formalism for field theory.
Consider now the electromagnetic field at $`x`$, emitted by this electron. $`\mathrm{\Delta }x=xz(\tau )`$ defines a family of four-vectors connecting the event $`x`$ to events on the electron worldline $`z(\tau ).`$ Then, accounting for the masslessness of the electromagnetic field, $`\mathrm{\Delta }x^2=0`$, the double lightcone with vertex at $`x,`$ intercepts $`z(\tau )`$ at two points: $`z(\tau _{ret})`$ and $`z(\tau _{adv}).`$ See the Figure 1.
The retarded field emitted by the electron at $`z(\tau _{ret})`$ must remain in the $`z(\tau _{ret})`$-future-lightcone, which contains x; and according to the standard interpretation , the advanced field produced by the electron at $`z(\tau _{adv})`$ must remain in the $`z(\tau _{adv})`$-past-lightcone, which also contains x. So, the electromagnetic field is defined just with local causality. There is then supposedly a clear dichotomy with respect to causality implementation in the treatment done to the electron and to its self-field , caused by the perception of the electron as a point particle, a discrete object, and of its field as a continuous and distributed one. Extended causality requires and implies discrete objects.
The (retarded and advanced) Liรจnard-Wiechert solutions of classical electrodynamics are
$$A^\mu (x)=\frac{eu^\mu (\tau )}{\rho }|_{\tau =\tau _s},\text{ for}\rho 0,$$
(5)
where $`\tau _s`$ stands for either $`\tau _{ret}`$ or $`\tau _{adv}`$, which are, respectively, the retarded and the advanced solution to the constraint
$$(xz(\tau ))^2=0,$$
(6)
imposed to $`A(x),`$ and
$$\rho :=u.\mathrm{\Delta }x,$$
with $`\mathrm{\Delta }x==xz(\tau )`$, represents $`|\mathrm{\Delta }\stackrel{}{x}|`$ in the charge rest-frame. Although $`A(x)`$ is restricted just by eq. (1), having thereby support on the lightcone, for the calculation of its Maxwell field,
$$F_{\mu \nu }:=_\nu A_\mu _\mu A_\nu ,$$
(7)
on a point $`x`$ it is necessary to consider $`A(x)`$ on a neighbourhood of $`x`$, and so a constraint equivalent to the eq. (3) must be also considered to assure the consistency of eq. (1) in this neighbourhood. From eq. (6) one has
$$\mathrm{\Delta }x.d(xz)=\mathrm{\Delta }x.(dxud\tau )=0,$$
that allows to write
$$d\tau +K.dx=0,$$
(8)
where K defined by
$$K^\mu =\frac{\mathrm{\Delta }x^\mu }{u.\mathrm{\Delta }x}|_{\tau _s}=\frac{\mathrm{\Delta }x^\mu }{\rho }|_{\tau _s},$$
(9)
is a null $`(K^2=0)`$ four-vector, tangent to the lightcone $`\mathrm{\Delta }x^2=0`$. $`K^\mu `$ shows the local direction of propagation of the electromagnetic field emitted by the electron at $`\tau _s.`$ In this context, eq. (8) is a consistency relation of eq. (2) assuring its validity for all successive pair of events $`(x,z(\tau )).`$ It implies on
$$K_\mu =\frac{\tau _s}{x^\mu },$$
(10)
and then,
$$\frac{1}{e}_\mu A^\nu |_{\tau _s}=\left\{\frac{K_\mu a^\nu }{\rho }+\frac{u^\nu }{\rho ^2}_\mu \rho \right\}|_{\tau _s}=\frac{1}{\rho ^2}\left(K_\mu W^\nu +u_\mu u^\nu \right)|_{\tau _s},$$
(11)
where $`a:=\frac{du}{d\tau },`$
$$_\mu \rho |_{\tau _s}=\{K_\mu (1+\rho a.K)u_\mu \}|_{\tau _s}$$
(12)
and the ancillary four-vector function $`W`$,
$$W^\mu =\{\rho a^\mu +u^\mu (1+\rho a.K)\}|_{\tau _s},$$
(13)
has been introduced just for notation simplicity. So,
$$F^{\mu \nu }=\frac{1}{\rho ^2}\left(K^\mu W^\nu K^\nu W^\mu \right)|_{\tau _s}$$
(14)
Geometrically the eq. (8), like the eq. (3), defines a family of hyperplanes that, for $`d\tau =0`$, are tangent to the lightcone (6), and parameterized by $`K_\mu =\eta _{\mu \nu }K^\nu .`$ The use of both constraints (6) and (8) implies the extended causality in the $`F`$ definition, exhibited on its explicit dependence on $`K^\mu `$, a four-vector tangent to a light-cone generator. But rigourously this is an inconsistent procedure as an undue mixing of local and extended causality on a same physical object. The inconsistency is on $`F`$ being defined as the curl of $`A(x)`$ which is a continuous field with support on the lightcone. If its support is reduced to the $`K`$-generator of its lightcone, $`F`$ should be regarded as a discrete object, similar, in this aspect, to its very source, the point electron. The problem, of course, is not with the definition (7) of $`F`$ but with $`A(x)`$ being a propagating field and, therefore, restricted by a causality constraint that necessarily requires the constraint (8) on any field derivative. In other words, there would be no problem with the definition (7) if $`A(x)`$ where not constrained by (6) as the constraint (8) would not be called up then<sup>*</sup><sup>*</sup>*A completely consistent formulation would require $`A(x)`$ being defined with extended causality too. This would imply the consideration of fields defined with support on (1+1) sub manifolds imbedded in the (3+1) spacetime, โdiscrete fieldsโ, with a complete symmetry between fields and sources, both being discrete objects. This is done in . The goal of the present paper is of pointing the existence of two modes for causality implementation (local and extended) in field theory and their implications to the meaning and nature of the fields and their interactions.
In the standard literature, without knowledge of extended causality, $`F`$ is seen as a field with support on the lightcone, i.e. a continuosly extended object with old and well-known problems with infinities and other inconsistencies. They just vanish after due consideration of extended causality .
The origin of this imbroglio is that the equation (8), as it can be formally obtained from a derivation of eq. (2), has been historically considered \[2-7\] as if all its effects were already described by eq. (2), included in it and not, as it is the case, a new and independent restriction to be considered at a same footing and in addition to it. An evidence of this is that eqs. (2) and (8) carry distinct physical informations that will be discussed now.
## IV Dynamics and causality
Eq. (8) connects the restriction on the propagation of the charge to the restriction on the propagation of its emitted or absorbed fields. Like its parent equation (2) it is just a kinematical restriction. But in the short-distance limit, when $`x`$ tends to $`z(\tau )`$, eq. (8), in contradistinction to eq. (2), is directly related to the changes in the charge state of movement due to the emission or to the absorption of electromagnetic field, that is, to the charge-field interaction process. Therefore, in this short distance limit eq. (8) also carries dynamical information, not only kinematical, as is the case of eqs. (1) and (2).
It is instructive to have a close look on the physical meaning of eqs. (6) and (8) for the case of an emitted field. Eq. (6) is a restriction on the propagation of a single object, the field emitted by the charge at $`z(\tau )`$, whereas the equation (8) connects restrictions on the propagation of two distinct physical objects, the electron and its field: $`d\tau `$ describes a displacement of the electron on its worldline while $`dx`$ is the four-vector separation between two other points where the electron self-field is being considered. If $`d\tau =0`$ then $`dx`$ is lightlike and collinear to K, as $`K.dx=0.`$ Thus, $`dx`$ is related to a same electromagnetic signal at two distinct times. The electromagnetic field at $`x+dx`$ can be seen as the same field at $`x`$ that has propagated to there with the speed of light. On the other hand, if $`d\tau 0`$ then $`dx`$ is not collinear to K and it is related to two distinct electromagnetic signals, emitted at distinct times. See the Figure 2. In this case, the field at $`x+dx`$ cannot be seen as the same field at $`x`$ that has propagated to there. It is another field emitted by the charge at another time. This apparently obvious interpretation of the constraint (8) reveals, however, deep physical implications as it perceives as being distinct objects the fields $`F`$ in two events that are not along a same $`K`$. This comes from extended causality requiring a $`F`$ defined with support on a $`K`$-light-cone generator and conflicting with local causality in the definition of $`A(x)`$.
But a $`F`$ defined with support on a lightcone generator produces strong and experimentally observable consequences. During the free propagation of an electromagnetic radiation, the four-vector $`K`$ of its light-cone-generator support must be constant. So, the eq. (8) implies on
$$1+K.u=0$$
(15)
and then
$$K.a=0.$$
(16)
The first one may be seen as a covariant normalization of $`K`$, that in the charge instantaneous rest frame must satisfy
$$K^0|_{\stackrel{}{u}=0}=|\stackrel{}{K}||_{\stackrel{}{u}=0}=1.$$
The second one is a dynamical constraint between the direction $`K`$ along which the signal is emitted (absorbed) and the instantaneous change in the charge state of motion at the retarded (advanced) time. It implies on
$$a_0=\frac{\stackrel{}{a}.\stackrel{}{K}}{K_0},$$
(17)
whereas $`a.u0`$ leads to $`a_0=\frac{\stackrel{}{a}.\stackrel{}{u}}{u_0},`$ and so, in the charge instantaneous rest frame at the limiting emission (absorption) time $`\stackrel{}{a}`$ and $`\stackrel{}{K}`$ are orthogonal vectors,
$$\stackrel{}{a}.\stackrel{}{K}|_{\stackrel{}{u}=0}=0.$$
(18)
This is an observable consequence of extended causality. For the electromagnetic field this is an old well known and experimentally confirmed fact. Its experimental confirmation is a direct validation of extended causality and of its implications, as discussed in the Section X of . The constraint (16) that takes, in the standard formalism of continuous fields, the whole apparatus of Maxwellโs theory to be demonstrated can been obtained on very generic grounds of causality without reference to any specific interaction. This makes of it a universal relation, supposedly valid for all kinds of fields and sources. This same behaviour, expressed in eq. (18), is then expected to hold for all fundamental (strong, weak, electromagnetic and gravitational) interactions.
The relevance of eq. (16) is on its focus on the charge-field interaction process. It is strongly dependent on $`K`$ being taken as a constant during the field propagation. A non-constant $`K`$ would imply on a continuing interaction and this would change the above results.
$$_\mu K_\nu =\frac{1}{\rho }(\eta _{\mu \nu }+K_\mu u_\nu +K_\nu u_\mu K_\mu K_\nu )K_\mu K_\nu a.K=_\nu K_\mu :=K_{\mu \nu },$$
(19)
from eqs. (9) and (12). Then the hypothesis of a non-constant $`K`$ would not affect eq. (15) because
$$K_{\mu \nu }\mathrm{\Delta }x^\nu =\rho K_{\mu \nu }K^\nu =K_\mu (1+K.u)0,$$
(20)
but eq. (16) would be replaced by just an identity as
$$_\mu (1+K_\nu u^\nu )=K_{\mu \nu }u^\nu K_\mu K_\nu a^\nu =K_\mu (1+K.u)0.$$
(21)
So, it is clear that the validity of eq. (16) rests on a free propagation of the field right after its emission (or, symmetrically, right before its absorption) which indicates no self-interaction, a definitive detachment of the field from its source. Self interaction for the emited field would also imply, by symmetry, causality violation for the absorbed field as it would be interacting with the charge even before reaching it.
## V The double limit $`xz(\tau )`$
The above conclusions can be made more evident considering the fate of both eqs. (2) and (8) in the limit when the event $`x`$ approaches the event $`z(\tau _s)`$ and its implications to the field energy-tensorThis is discussed in but for completeness, considering its relevance here, its main steps and some further considerations are aligned.. Nothing obviously happens to the first one; $`\mathrm{\Delta }x`$ just goes to zero. To the second one the restriction connecting $`d\tau `$ to $`dx`$ becomes indeterminated because K is not defined in this limit:
$$\underset{xz(\tau _s)}{lim}K=\underset{xz(\tau _s)}{lim}\frac{\mathrm{\Delta }x}{u.\mathrm{\Delta }x}=\frac{0}{0}\mathrm{?}$$
(22)
For a lightlike signal, eqs. (2) and (8) together require that the pair of events $`x`$ and $`z(\tau )`$ belongs to a same lightcone generator, so that eq. (22) can be written as
$$\underset{xz(\tau _s)}{lim}K|_{\genfrac{}{}{0pt}{}{\mathrm{\Delta }x^2=0}{d\tau +K.dx=0}}=\frac{0}{0}\mathrm{?}$$
(23)
This notation intends to denote that $`x`$ approaches $`z(\tau _s)`$ through a K-light-cone generator, i.e. by the straight line intersection of the hypercone $`(\mathrm{\Delta }x^2=0`$) and its tangent hyperplane ($`d\tau +K.dx=0`$), eliminating any ambiguity in the definition of the limit in eq. (22). Now one can apply the LโHรดpitalโs rule for evaluating K on the neighbouring events of $`z(\tau _s)`$ along the electron worldline, i.e., at either $`\tau _s+d\tau `$ or $`\tau _sd\tau .`$ This corresponds to replacing the above simple limit of $`xz(\tau _s)`$ by a double and simultaneous limit of $`xz(\tau )`$ along the K-lightcone generator while $`z(\tau )z(\tau _s)`$ along the electron worldline. This simultaneous double limit is pictorially best described by the sequence of points S, Q,โฆ,P in the Figure 3; each point in this sequence belongs to a K-generator of a lightcone with vertex at the electron worldline $`z(\tau ).`$
Then, from eq. (22),
$$\underset{\genfrac{}{}{0pt}{}{xz(\tau )}{\tau \tau _s}}{lim}K|_{\genfrac{}{}{0pt}{}{\mathrm{\Delta }x^2=0}{d\tau +K.dx=0}}=\underset{\genfrac{}{}{0pt}{}{xz(\tau )}{\tau \tau _s}}{lim}\frac{\mathrm{\Delta }\dot{x}}{a.\mathrm{\Delta }x+u.\mathrm{\Delta }\dot{x})}|_{\genfrac{}{}{0pt}{}{\mathrm{\Delta }x^2=0}{d\tau +K.dx=0}}=$$
$$=\underset{\genfrac{}{}{0pt}{}{xz(\tau )}{\tau \tau _s}}{lim}\frac{u}{u.u}|_{\genfrac{}{}{0pt}{}{\mathrm{\Delta }x^2=0}{d\tau +K.dx=0}}=u,$$
(24)
as $`\dot{\mathrm{\Delta }}x:=\frac{d\mathrm{\Delta }x}{d\tau }=u`$ and $`u^2=1.`$ So $`K|_{x=z(\tau _s)}`$ is indefinite but $`K|_{x=z(\tau _s\pm d\tau )}=u.`$
The lightlike four-vector K is replaced by the timelike four-vector u in the above defined (double) limit of $`\mathrm{\Delta }x0.`$ This result changes the usual vision of field theory in this limit.
The electron self-field energy tensor, $`4\pi \mathrm{\Theta }=F.F\frac{\eta }{4}F^2`$, after eq. (11) becomes,
$$4\pi \rho ^4\mathrm{\Theta }=(KW+WK)+KKW^2+WWK^2+\frac{\eta }{2}(1K^2W^2),$$
(25)
as $`K.u=1`$ from eq. (9) and $`K.W=1`$. The eq. (25), like eqs. (5) and (14), are all constrained by eq. (2), i.e. by $`\tau =\tau _s,`$ and they are valid only for $`\rho 0`$, region where $`K^2=0.`$ So, instead of eq. (25) one may write
$$4\pi \rho ^4\mathrm{\Theta }|_{K^2=0}=(KW+WK)+KKW^2+\frac{\eta }{2},\text{ for}\rho >0,\tau =\tau _s,$$
(26)
which corresponds to the usual expressions that one finds, for example in \[2-7\]. They are equivalent, as long as $`\rho >0`$. The four-vector momentum associated to the electron self-field is defined by the flux of its $`\mathrm{\Theta }`$ through a hypersurface $`\sigma `$ of normal n:
$$P=d^3\sigma n.\mathrm{\Theta }|_{K^2=0},$$
(27)
but $`\mathrm{\Theta }`$ contains a factor $`\frac{1}{(\rho )^4}`$ and this makes P highly singular at $`\rho =0`$, that is at $`x=z(\tau _s).`$ This is the old well-known self-energy problem of classical electrodynamics which heralds similar problems in its quantum version. This divergence at $`\rho =0`$ is also the origin of nagging problems on finding a classical equation of motion for the electron \[2-7\]. But it is clear now, after equation (24), that the standard practice of replacing everywhere $`\mathrm{\Theta }`$ by $`\mathrm{\Theta }|_{K^2=0}`$ is not justified and, more than that, it is the cause of the above divergence problem and the related misconceptions in classical electrodynamics. One must use eq. (25), the complete expression of $`\mathrm{\Theta }`$, in eq. (27) and repeat for it the same double limit done in eq. (24). The long but complete and explicit calculation is done in ; its results and conclusions are summarised here:
$`P|_{x=z(\tau _s)}`$ is undefined but
$$P|_{x=z(\tau _s)}=P|_{x=z(\tau _s+)}=0.$$
(28)
There is no infinity at $`\rho =0`$! This infinity disappears only when the double limiting process is taken because the lightcone generator $`K`$ must then be recognized as the actual support of the Maxwell field $`F.`$ The message here is that the infinities and other inconsistencies of classical electrodynamics are not to be blamed on the point electron but on the lightcone support of the field in the eq. (5). Extended causality cannot just be ignored.
## VI Conclusions
We can summarize it all with the following implications to the charge-field dynamics:
1. No self interaction; once emitted the field no longer interacts with the charge;
2. The emission process is discrete;
3. The emission event is an isolated singularity on the charge worldline, in the sense of discontinuity on its first derivative.
Equation (28) confirms that $`z(\tau _s)`$ is an isolated singularity. This is in direct contradiction to the standard view of a continuous field, emitted or absorbed by the charge in a continuous way. According to eq. (28) there is no charge self field at $`z(\tau _s\pm d\tau )`$, but only sharply at $`z(\tau _s)`$. This is puzzling! It is saying that the Gaussโ law, in the zero distance limit, $`lim_{S0}_S๐\sigma \stackrel{}{E}.\stackrel{}{n}=4\pi e`$, is meaningful only at $`z(\tau _s)`$ and not at $`z(\tau _{ret\pm d\tau })`$ because $`\stackrel{}{E}(\tau _s)0`$ but $`\stackrel{}{E}(\tau _{ret\pm d\tau })=0.`$
It implies, in other words, that the electromagnetic interactions are discrete and localized in time and in space. In terms of a discrete field interaction along a lightcone generator, as the one representedFor simplicity, the cause(the absorption of a photon, for example) of the sudden change in the electron state of movement is not shown, only its consequence (the emission of a photon) in the Figure 4, one can understand the physical meaning of eqs. (23), (24) and (28). The Maxwell fields are just effective average descriptions of an actually discrete interaction field. The field discreteness (or the existence of photons) is masqueraded by this averaged field and it takes the zero distance limit to be revealed from the Maxwell field. It may sound unbelievable or even suspicious that these conclusions have been derived exclusively from the supposedly exhaustively known classical electrodynamics but nothing has been added to or modified in the old Maxwellโs theory, except a new interpretation of old results. They come from the recognition of the existence of two mutually excludent causality-implementation modes in the Maxwellโs formalism. This could have been taken, if it were known at the beginning of the last century, as a first indication of the quantum, or of the discrete nature of the electromagnetic interaction. All these are consequences of the dynamical constraints hidden on the restrictions (2) and (8).
The initial goal of pointing and discussing the implicit existence of two distinct modes of implementing causality in field theory and that its consistency problems are created by non-recognizing them has been fulfilled. A completely consistent field formalism must be expressed in term of fields defined ab initio with extended causality. How this can be accomplished, its consequences and how it is related to the standard formalism based on local causality only is done in .
|
warning/0006/hep-ph0006205.html
|
ar5iv
|
text
|
# Quark-Gluon Plasma as a Condensate of ๐โข(3) Wilson Lines
## Abstract
Effective theories for the thermal Wilson line are constructed in an $`SU(N)`$ gauge theory at nonzero temperature. I propose that the order of the deconfining phase transition for $`Z(N)`$ Wilson lines is governed by the behavior of $`SU(N)`$ Wilson lines. In a mean field theory, the free energy in the deconfined phase is controlled by the condensate for $`Z(N)`$ Wilson lines. Numerical simulations on the lattice, and the mean field theory for $`Z(3)`$ Wilson lines, suggest that about any finite temperature transition in QCD, the dominant correlation length increases by a large, uniform factor, of order five.
A new phase of matter, the Quark-Gluon Plasma, might be produced in the collisions of large nuclei at very high energies. By asymptotic freedom, the pressure approaches the ideal gas value in the limit of high temperature, and so it is natural to think of the high temperature phase of QCD as a gas of quasiparticles .
It is known, however, that the high temperature phase of a purely glue theory is like the low temperature phase of a spin system. The magnetization in the high temperature phase of a $`SU(N)`$ gauge theory is a $`Z(N)`$ spin, proportional to the trace of the thermal Wilson line .
In this paper I construct effective lagrangians for the thermal Wilson line, considered as a full $`SU(N)`$ matrix, as well as its trace. This leads to novel sigma models of adjoint $`SU(N)`$ fields. Although the critical behavior is inexorably governed by the fixed point of $`Z(N)`$ spins , the $`SU(N)`$ spins can be important. In particular, they help explain why the order of the deconfining transition appears to change with $`N`$: from second order for $`N=2`$ , to weakly first order for $`N=3`$ , to first order for $`N4`$ . Further, the picture of the high temperature phase is turned on its head: the pressure isnโt due to quasiparticles , but is a potential for a condensate of $`Z(N)`$ Wilson lines. A mean field theory then suggests that because the deconfining transition in pure glue $`SU(3)`$ is weakly first order, QCD is near a critical point. About the transition, the dominant correlation lengths increase by a large factor, of order five .
I concentrate on the pure glue theory; later I argue why this is legitimate, using the lattice data and the effective theory. The thermal Wilson line is
$$๐(x)=๐ซ\mathrm{exp}\left(ig_0^{1/T}A_0(x,\tau )๐\tau \right),$$
(1)
where $`๐ซ`$ is path ordering, $`g`$ is the gauge coupling constant, $`A_0`$ is the time component of the vector potential in the fundamental representation, $`x`$ is the coordinate for three spatial dimensions, and $`\tau `$ that for euclidean time at a temperature $`T`$. The Wilson line in (1) is a product of $`SU(N)`$ matrices, and so is itself a $`SU(N)`$ matrix, satisfying
$$๐^{}(x)๐(x)=\mathrm{๐},det(๐(x))=1$$
(2)
Without quarks, the allowed gauge transformations are periodic up to an element of a global $`Z(N)`$ symmetry :
$$๐(x)\mathrm{exp}(2\pi i/N)\mathrm{\Omega }^{}(x)๐(x)\mathrm{\Omega }(x),$$
(3)
$`\mathrm{\Omega }(x)=\mathrm{\Omega }(x,0)`$. $`๐(x)`$ transforms as an adjoint field under local $`SU(N)`$ gauge transformations in three dimensions, and as a vector under global $`Z(N)`$ transformations.
Effective theories for $`๐(x)`$ are dictated by the symmetries of (3). I begin with the nonlinear form, where (2) are taken as constraints on $`๐`$. I construct an effective theory in three spatial dimensions, valid for distances $`1/T`$, by coupling the gauge potentials for static magnetic fields, the $`A_i(x)`$โs, to the Wilson line, $`๐(x)`$:
$$_0=\frac{1}{2}\mathrm{tr}\left(G_{ij}^2\right)+T^2a_1\mathrm{tr}|D_i๐|^2,a_1=\frac{1}{g^2}+\mathrm{}$$
(4)
The first term is the standard lagrangian for static $`A_i`$ fields (by choice, $`A_i`$ has dimensions of mass; all lagrangians have dimensions of (mass)<sup>4</sup>). In the second term, I start with electric part of the gauge lagrangian, $`\mathrm{tr}|D_iA_0|^2`$, and assume that it transmutes into a gauge invariant kinetic term for $`๐(x)`$. This is the continuum form of the lattice model of Banks and Ukawa .
Notice the factor of $`T^2`$ in front of the kinetic term for $`๐`$. This arises because the Wilson line is a phase in color space, and so every element is a dimensionless pure number. Thus in any effective lagrangian, dimensions can only be made up by powers of the temperature $`T`$.
Consider the somewhat peculiar limit in which one drops the coupling to the $`A_i`$โs, by taking $`g0`$, but retains $`๐\mathrm{๐}`$. Then (4) reduces a nonlinear sigma model in three dimensions, with lagrangian $`\mathrm{tr}|_i๐|^2`$. With the constraints of (2), the theory is invariant under $`๐\mathrm{\Omega }_1๐\mathrm{\Omega }_2`$, where $`\mathrm{\Omega }_1`$ and $`\mathrm{\Omega }_2`$ are independent, constant $`SU(N)`$ matrices. This is an enhanced global symmetry of $`SU(N)\times SU(N)`$ (times the usual global $`Z(N)`$ symmetry). As a sigma model, it is possible to impose other constraints upon $`๐`$ beyond those of (2). For example, requiring $`\mathrm{tr}๐`$ to be some fixed number produces a sigma model on a symmetric space .
At nonzero coupling, $`๐`$ is simply an adjoint field under the local $`SU(N)`$ symmetry. Even with the constraints of (2), the reduced symmetry implies that many more terms arise: instead of $`\mathrm{tr}๐`$ being fixed, as for a symmetric space, arbitrary traces, such as $`\mathrm{tr}๐^p`$ for integer $`p`$, are allowed. Mathematically, $`\mathrm{tr}๐^p`$ is related to the trace of the Wilson line in higher representations .
At one loop order, the terms up to fourth order in $`A_0`$ have been computed. The quadratic term, $`\mathrm{tr}(A_0^2)`$ , is the Debye mass for the gluon. For $`N4`$, there are two independent quartic terms, $`(\mathrm{tr}(A_0^2))^2`$ and $`\mathrm{tr}(A_0^4)`$ , which represent a potential for $`A_0`$. From (1), it is easy to turn a potential for $`A_0`$ into one for $`๐`$:
$$_1=T^4\left(c_2|\mathrm{tr}๐|^2+c_4|\mathrm{tr}๐^2|^2+c_4^{}\left(|\mathrm{tr}๐|^2\right)^2\right).$$
(5)
Expanding to $`A_0^4`$ fixes $`c_2=(4+3/\pi ^2)/9`$, $`c_4=+(1+3/\pi ^2)/36`$, and $`c_4^{}=0`$. The rational terms in $`c_2`$ and $`c_4`$ are from the Debye mass, while those $`1/\pi ^2`$ arise from quartic terms in the potential for $`A_0`$. Only two constants, $`c_2`$ and $`c_4`$, are needed to fit three terms in the $`A_0`$ potential. If $`N_f`$ flavors of massless quarks are included, $`c_2`$ and $`c_4`$ change, while $`c_4^{}`$ is then nonzero.
The signs of $`c_2`$ and $`c_4`$ are interesting. As $`๐g^2A_0^2`$, a positive Debye mass corresponds to negative $`c_2`$. The coupling $`c_4`$ is like the quartic term in $`A_0`$, and so positive. Negative $`c_2`$ favors condensation in a direction in which $`|\mathrm{tr}๐|^2`$ is maximized. This happens when $`๐`$ is an element of the center ,
$$๐=\mathrm{exp}(2\pi ij/N)\mathrm{}_0\mathbf{\hspace{0.33em}1},$$
(6)
$`j=0\mathrm{}(N1)`$. Different $`j`$ are the usual $`N`$ degenerate vacua of the broken $`Z(N)`$ global symmetry.
In (6) I introduce an expectation value, $`\mathrm{}_0=\mathrm{}`$, where $`\mathrm{}`$ is defined in (8). In perturbation theory, $`\mathrm{}_0=1`$, but $`\mathrm{}_0`$ is a function of temperature; it vanishes at the critical temperature, $`T_c`$, and in the confined phase, for $`T<T_c`$.
I assume that in the deconfined phase, $`T>T_c`$, the stable vacuum is that which maximizes $`|\mathrm{tr}๐|^2`$, so that $`๐`$ condenses as in (6). An expectation value for a field in the fundamental representation always breaks the gauge symmetry, but uniquely for an adjoint field, a vacuum expectation value proportional to the unit matrix does not: (6) is invariant under arbitrary local gauge rotations. Similarly, the adjoint covariant derivative in (4) is $`D_i๐=_i๐ig[A_i,๐]`$, so with (6), the static magnetic gluons do not acquire a mass when $`๐`$ condenses, $`\mathrm{}_00`$. Thus (6) is the nonperturbative statement that electric screening does not generate screening for static magnetic fields .
The terms in (5) are invariant under a global symmetry of $`U(1)`$. There are also terms which reduce this $`U(1)`$ to $`Z(N)`$. For $`N=3`$, the simplest examples include
$$det๐+\mathrm{c}.\mathrm{c}.,(\mathrm{tr}๐)^3+\mathrm{c}.\mathrm{c}.,\mathrm{tr}๐(\mathrm{tr}๐^2)+\mathrm{c}.\mathrm{c}..$$
(7)
The first term, $`det๐`$, is $`SU(3)\times SU(3)`$ symmetric, while the others are only $`SU(3)`$ symmetric.
There are also a wide variety of kinetic terms possible. These include $`|_i\mathrm{tr}๐|^2`$, $`|_i\mathrm{tr}๐^2|^2`$, and $`|\mathrm{tr}๐|^2\mathrm{tr}|D_i๐|^2`$, amongst others. At one loop order, the kinetic term in (3) is renormalized, and terms such as these may be generated; present calculations cannot distinguish . Even for $`g=0`$, none of these new kinetic terms are invariant under $`SU(N)\times SU(N)`$.
The potential in (5) is only illustrative. In perturbation theory, one expands about $`๐\mathrm{๐}`$, which does not allow one to uniquely fix the coefficients of a potential for $`๐`$. Through numerical simulations, effective theories for $`A_0`$ , and those for $`๐`$, could be matched by comparing physical correlation lengths at an intermediate temperature scale, say at several times the critical temperature.
Now consider a point of second order transition, where $`\mathrm{}_0(T)0`$. Then powers of $`๐`$ are suppressed, and it is sensible to construct a linear sigma model. This is done by introducing an โblock spinโ $`๐`$, formed by a gauge covariant average of $`๐`$ over some region of space . Any $`SU(N)`$ matrix can be written as
$$๐(x)=\mathrm{}(x)\mathrm{๐}+2i\stackrel{~}{\mathrm{}}_a(x)t^a,$$
(8)
where $`t^a`$ are the generators of $`SU(N)`$, $`a=1\mathrm{}(N^21)`$. For general $`N`$, $`\mathrm{}`$ and $`\stackrel{~}{\mathrm{}}_a`$ are complex valued, and (2) imposes $`N^2+1`$ constraints.
I start with the case of two colors, which is special. Four constraints of (2) are satisfied in an especially simple manner: the imaginary parts of $`\mathrm{}`$ and $`\stackrel{~}{\mathrm{}}_a`$ vanish. This leaves one constraint, which is $`\mathrm{}^2+\stackrel{~}{\mathrm{}}_a^2=1`$; thus $`\mathrm{}`$ and $`\stackrel{~}{\mathrm{}}_a`$ form a vector representation of $`SU(2)\times SU(2)=O(4)`$. After averaging, the constraint on the $`O(4)`$ norm is lost, as is typical in a linear model. Averaging still leaves $`\mathrm{}`$ and $`\stackrel{~}{\mathrm{}}_a`$ as real valued fields, though. Up to quartic order, the most general lagrangian is
$$=\frac{1}{2}\mathrm{tr}\left(G_{ij}^2\right)+\frac{1}{2}(_i\mathrm{})^2+\mathrm{tr}|D_i\stackrel{~}{\mathrm{}}|^2$$
(9)
$$m_1(\mathrm{}^2+\stackrel{~}{\mathrm{}}_a^2)m_2\mathrm{}^2+\lambda _1(\mathrm{}^2+\stackrel{~}{\mathrm{}}_a^2)^2+\lambda _2\mathrm{}^4+\lambda _3\mathrm{}^2\stackrel{~}{\mathrm{}}_a^2.$$
The $`\mathrm{}`$-field is a color singlet, while $`\stackrel{~}{\mathrm{}}_a`$ is an adjoint $`SU(2)`$ field. The terms $`m_1`$ and $`\lambda _1`$ are $`O(4)`$ symmetric; with the kinetic terms, they correspond to the gauged nonlinear sigma model of (4). The other terms, $`m_2`$, $`\lambda _2`$ and $`\lambda _3`$, correspond to the potential for the Wilson line in (5). A factor of $`T`$ has been absorbed into the definition of $`\mathrm{}`$ and $`\stackrel{~}{\mathrm{}}_a`$.
When $`N3`$, the constraints of (2) are nonlinear. Since a sum of two special unitary matrices is not necessarily special unitary, the average $`๐`$ must be taken to be a complex $`N\times N`$ matrix. Thus $`๐`$ includes a complex valued, color singlet field, $`\mathrm{}`$, which I call a $`Z(N)`$ spin, and a complex valued, color adjoint field, $`\stackrel{~}{\mathrm{}}_a`$, which I call a $`SU(N)`$ spin.
Linear models like (9) can be written down for $`N3`$, although there is a plethora of terms. At quartic order there is one term which is $`O(2N^2)`$ symmetric, $`(\mathrm{tr}๐^{}๐)^2`$, another which is $`SU(N)\times SU(N)`$ symmetric, $`\mathrm{tr}(๐^{}๐)^2`$, and terms which are only invariant under $`SU(N)`$, such as ($`\stackrel{~}{\mathrm{}}\stackrel{~}{\mathrm{}}_at_a`$)
$$(|\mathrm{}|^2)^2,\mathrm{}\mathrm{tr}(\stackrel{~}{\mathrm{}}^{})^2\mathrm{}+\mathrm{c}.\mathrm{c}.,\mathrm{}^2\mathrm{tr}(\stackrel{~}{\mathrm{}}^{})^2+\mathrm{c}.\mathrm{c}.,$$
$$(\mathrm{tr}\stackrel{~}{\mathrm{}}^{}\stackrel{~}{\mathrm{}})^2,|\mathrm{tr}\stackrel{~}{\mathrm{}}^2|^2,\mathrm{tr}(\stackrel{~}{\mathrm{}}^{}\stackrel{~}{\mathrm{}})^2,\mathrm{tr}(\stackrel{~}{\mathrm{}}^{})^2\stackrel{~}{\mathrm{}}^2.$$
(10)
These models give a qualitative picture of the deconfining phase transition: I assume that while only the $`Z(N)`$ $`\mathrm{}`$-spins condense, $`\mathrm{}\mathrm{}_00`$, that the transition is driven by the behavior of the $`SU(N)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins. This picture is based on the nonlinear model: at weak coupling, (4) dominates other terms, such as (5), by $`1/g^2`$. Now certainly all coupling constants change with $`T`$, as can be seen from the temperature dependence of the Debye mass . Nevertheless, I assume that the $`SU(N)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins dominate right down to the point of the deconfining phase transition. The only purpose of terms such as (5) is to ensure that condensation which respects the local $`SU(N)`$ symmetry, (6), is favored.
For two colors, the influence of the $`SU(2)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins on the $`Z(2)`$ $`\mathrm{}`$-spins is subtle. Assume that only the $`O(4)`$ symmetric mass, $`m_1`$, changes. The phase transition in a gauged $`SU(2)`$ model is known from lattice studies of the electroweak phase transition . I assume that one is always in an extreme โtype-IIโ regime, so that the second order $`O(4)`$ transition of the model with $`g=0`$ (the point $`B_2`$ of fig. (1) in ) is washed out by confinement of nonabelian gauge fields. The only transition is a point at which the $`Z(2)`$ $`\mathrm{}`$-spins become massless; the $`SU(2)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins are always massive. Lattice studies confirm a second order transition in the $`Z(2)`$ universality class .
For $`N3`$, the $`SU(N)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins can have first order transitions. This is because in the absence of gauge fields, $`SU(N)\times SU(N)`$ spin models have first order transitions both in mean field theory and in an expansion about $`4ฯต`$ dimensions . As suggested in , in the extreme type-II regime, confinement of the gauge fields need not wash out the first order transition of $`SU(N)\times SU(N)`$ spins (above the point $`B_3`$ in fig. (2) of ), and so the deconfining transition can remain first order. In particular, the transition can be of first order as $`N\mathrm{}`$. This is in accord with a lattice analysis of Gocksch and Neri , and contrary to previous speculation .
For three colors, this implies that the deconfining transition is of first order not only because of cubic invariants , as in (7), but because of the dynamics of $`SU(3)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins. Relative to the ideal gas, the latent heat for three colors is $`1/3`$ . As could have been guessed from the lattice data alone, perhaps the deconfining transition is weakly first order for $`N=3`$ because it is near the second order transition for $`N=2`$. Thus it is of value to know how the latent heat for $`N=4`$ compares to that for $`N=3`$: is it more strongly first order, such as $`(N2)/N`$ as $`N2`$, or more weakly first order, like $`1/N`$ as $`N\mathrm{}`$ ?
Whatever the order of the deconfining phase transition, one can write a mean field theory in which the free energy in the deconfined phase is controlled by a potential for the $`Z(N)`$ Wilson lines. For three colors, this is :
$$๐ฑ=\left(2b_2|\mathrm{}|^2+b_3(\mathrm{}^3+(\mathrm{}^{})^3)+(|\mathrm{}|^2)^2\right)b_4T^4.$$
(11)
$`\mathrm{}`$ is complex valued, so when $`b_30`$, the global symmetry is reduced from $`O(2)`$ to $`Z(3)`$. The coupling $`b_3`$ must be small for the transition to be weakly first order , so for now I ignore it, considering the potential just as a function of $`b_2`$ and $`b_4`$. This is similar to the case of two colors, where $`\mathrm{}`$ is a real field, and the potential is just a sum of two terms, $`b_2\mathrm{}^2`$ and $`b_4\mathrm{}^4`$ .
In speaking of the Wilson line, implicitly I assume that it is possible to extract a renormalized value from the bare quantity . If so, then given $`\mathrm{}_0(T)`$ and the pressure, one could fit to a potential like (11); for example, is it necessary to include higher powers of $`\mathrm{}`$ in $`๐ฑ`$?
The novel aspect of (11) is my insistence that because $`\mathrm{}`$ is a dimensionless field, the dimensions in $`๐ฑ`$ must be made up by the temperature, $`T`$. In mean field theory, $`b_4`$ is taken as constant, and $`b_2`$ varies with temperature, vanishing at $`T_c`$. The pressure is given by the minimum of the potential, $`p=b_2^2b_4T^4`$, and vanishes in the confined phase, $`T<T_c`$. That is, with the overall $`T^4`$ in the potential, the pressure is like a gas of quasiparticles, albeit with a variable number of degrees of freedom, which vanish at $`T_c`$.
At high temperature, $`b_21`$ so that $`\mathrm{}_01`$. The quartic coupling is fixed by the ideal gas limit: if $`n_{\mathrm{}}=p/T^4`$ as $`T\mathrm{}`$, $`b_4=n_{\mathrm{}}`$. Lattice simulations find that the $`p/T^4`$ is relatively flat down to a scale which is several times $`T_c`$, call it $`\kappa T_c`$; the same is found from resummations of perturbation theory . ($`\kappa `$ might be defined as the lowest value where $`\mathrm{}_01`$.) Hence I assume that $`b_2`$ and $`b_4`$ are slowly varying down to $`\kappa T_c`$.
Between $`\kappa T_c`$ and $`T_c`$, I assume that $`b_4`$ is essentially constant, while $`b_2`$ varies. In particular, the trace of the energy momentum tensor, divided by $`T^4`$, is $`(e3p)/T^4=T(b_2^2b_4)/T`$. Lattice simulations find that this quantity has a peak just above $`T_c`$ : this is then due to the rapid variation of $`b_2`$ with temperature .
In the $`Z(N)`$ mean field theory, the pressure includes only the contribution of the potential, and nothing from fluctuations in the effective fields, either from the $`Z(N)`$ $`\mathrm{}`$-spins or the $`SU(N)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins. Fluctuations in these fields do, of course, contribute to the pressure at all temperatures. Since by construction the pressure in the mean field approximation vanishes for $`T<T_c`$, one condition for its validity is that the pressure in the confined phase is small. This is what present lattice simulations find. Physically, these fields donโt contribute much to the pressure because they are heavy: the $`SU(N)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins always so, and the $`Z(N)`$ $`\mathrm{}`$-spins usually so. For two or three colors, the $`\mathrm{}`$-spins do become light in a narrow band in temperature about $`T_c`$, where mean field theory fails .
Ignoring fluctuations about the mean field theory is also justified from the viewpoint of an expansion in a large number of colors, $`N\mathrm{}`$ . The free energy in the confined phase is of order one, while it is $`N^2`$ in the deconfined phase. The term $`N^2`$ in the free energy is due entirely to the condensate, taking $`b_21`$ and $`b_4N^2`$. Even though there are $`N^2`$ of them, the $`SU(N)`$ $`\stackrel{~}{\mathrm{}}_a`$-spins only contribute to the pressure at $`1`$, since they are bound into color singlet glueballs. The $`Z(N)`$ $`\mathrm{}`$-spins also contribute $`1`$ to the free energy.
I have concentrated on the pure glue theory because numerical simulations have demonstrated the following remarkable property . If $`p/(n_{\mathrm{}}T^4)`$ is plotted versus $`T/T_c`$, the resulting curve is nearly universal, and looks very similar whether or not there are dynamical quarks present. The present model predicts that the pressure is the same because the (renormalized ) Wilson line is the same. In terms of the potential, (11), the differences in the ideal gas values, $`n_{\mathrm{}}`$, are absorbed into $`b_4`$, with the same $`b_2(T/T_c)`$.
Quarks act like a background magnetic field for the real part of $`\mathrm{}`$ . Because the pure glue transition is weakly first order, it is not difficult for quarks to wipe out the deconfining transition altogether, leaving either a chiral transition, or just crossover behavior. Even so, what is relevant here is that for three colors and two or three flavors of quarks, the pressure for $`T<T_c`$ is always much smaller than that for $`T>T_c`$; that is, up, down, and strange quarks act like a weak magnetic field for the $`Z(3)`$ $`\mathrm{}`$-spins.
I thus come to the central physical point of this paper. The lattice tells us that the deconfining transition in pure glue $`SU(3)`$ theory is close to the second order transition for $`SU(2)`$; further, that the effects of quarks are small, except close to $`T_c`$. I suggest that what is important is not whether the weakly first order transition persists with quarks, but that the nearly second order transition very well might. In the pure glue theory, as $`TT_c^+`$ the ratio of the screening mass to the temperature decreases by a factor of ten: from $`2.5`$ at $`T2T_c`$, to $`.25`$ at $`TT_c^+`$ . Similarly, the string tension at $`TT_c^{}`$ is ten times smaller than that at zero temperature . With quarks, the increase in the correlation length for $`\mathrm{}`$ is presumably less, maybe not ten, but perhaps a factor of five or so. And most importantly, if the pressure below $`T_c`$ is small, it might be justified to use the $`Z(3)`$ mean field theory.
If the chiral order parameter is $`\mathrm{\Psi }`$, then it couples to $`Z(3)`$ $`\mathrm{}`$-spins through the coupling $`+|\mathrm{}|^2\mathrm{tr}\left(\mathrm{\Psi }^{}\mathrm{\Psi }\right).`$ Lattice simulations find that the chiral and deconfining transitions occur at approximately the same temperature. This naturally results if this coupling constant is positive, as condensation in one field tends to suppress condensation in the other. Coherent oscillations in the $`\mathrm{}`$-field couple to light mesons through such a term, and can produce large fluctuations in the average pion momentum .
This uniform increase in correlation lengths near $`T_c`$ is a unique prediction of the $`Z(3)`$ mean field theory. In quasiparticle models of the quark-gluon plasma, the pressure is tuned to vanish at $`T_c`$ by the introduction of a bag constant. In order for the energy to decrease as $`TT_c^+`$, though, the quasiparticles must become heavier, not lighter; that is, instead of increasing, most correlation lengths decrease .
At nonzero quark chemical potential, $`\mu `$, presumably there is little change if the quarks are hot and dilute: for small $`\mu /T`$, the $`Z(3)`$ $`\mathrm{}`$-spins should still exhibit nearly second order behavior. I contrast this with the (possible) critical endpoint of the chiral transition in the $`\mu T`$ plane . The correlation length of the sigma meson truly diverges at the critical endpoint, but this only occurs at one special value of $`\mu `$. Moreover, the sigma meson does not dominate the free energy, nor generic particle production. For cold, dense quark matter, $`\mu T`$, I do not see why $`Z(3)`$ $`\mathrm{}`$-spins should dominate the free energy.
I conclude by noting that the generalization of the Debye mass term, $`\mathrm{tr}A_0^2`$, to real scattering processes produces hard thermal loops . This is then the first term in an infinite series of such terms, continuing $`\mathrm{tr}A_0^4`$, etc. The natural expansion is not in powers of $`A_0`$, but in powers of the Wilson line, as in (5). It is then of great interest to know the analytic continuation of the Wilson line to real scattering processes .
I benefited from discussions with K. Eskola, F. Gelis, K. Kajantie, F. Karsch, C. P. Korthals-Altes, M. Laine, D. Miller, S. Ohta, A. Peshier, D. H. Rischke, M. Wingate (for the argument in ), L. Yaffe, and, especially, M. Creutz. This work is supported by DOE grant DE-AC02-76CH00016.
|
warning/0006/astro-ph0006397.html
|
ar5iv
|
text
|
# MASS TO LIGHT RATIOS OF GROUPS AND CLUSTERS OF GALAXIES
## 1. INTRODUCTION
Galaxy clusters are the most massive bound systems known and hence are of interest for investigating cosmological parameters. These can be constrained by studying fundamental properties such as cluster mass-to-light ratios, dark matter distributions, gas mass fractions or the ratios of luminous baryon mass to the total mass.
The cosmological parameter $`\mathrm{\Omega }_0`$ (the ratio of the mass density of the universe to the critical density) can be constrained by measuring the mass-to-light ratios of clusters, estimating the luminosity density of the universe, and assuming that clusters have a dark matter content representative of the whole Universe. This assumption is supported by measurements of the Virgo cluster infall motion, the cosmic virial theorem (Bahcall et al. 1995) and a weak gravitational lensing mass estimate of a supercluster of galaxies, yielding a mass-to-light ratio comparable to that of clusters, $`\mathrm{M}/\mathrm{L}=(140\pm 20)h_{50}\mathrm{M}_{}/\mathrm{L}_{}`$ (Kaiser et al. 1998)
$`\mathrm{\Omega }_0`$ also can be independently constrained by studying the cluster gas mass fractions (the ratio of the gas mass to total mass). Predictions from standard big bang nucleosynthesis limit the baryon density of the universe to $`\mathrm{\Omega }_b=f_b\mathrm{\Omega }_0=0.076\pm 0.004h_{50}^2`$ (Walker et al. 1991, White et al. 1993, Tytler et al. 1996, Kirkman et al. 2000), where $`f_b`$ is the baryon mass fraction. The luminous baryonic component of clusters consists primarily of the intra-cluster gas, with a small contribution from stars. The possible other components not observed result in the luminous baryons being a lower limit on the baryon fraction. Assuming that the observed baryon fractions in clusters are representative of the baryonic content of the whole universe, we can use cluster gas mass fractions to place an upper limit on $`\mathrm{\Omega }_0`$, given the Hubble constant $`H_0`$.
These investigations fundamentally rely on accurate measurements of cluster luminosities and masses. Summed optical luminosities of clusters are ideally measured using CCD observations. However, photometric data sets extending to the cluster virial radii are available only for relatively small samples for low redshift clusters (Lopez-Cruz 1995). A presently feasible way of measuring low redshift cluster luminosities over large volumes is to use photometric CCD images in a suitable filter system to calibrate photographic survey plates covering larger areas of the sky. We use the Digitized Second Palomar Sky Survey photographic plates (Djorgovski et al. 1998), calibrated with CCD images in the Gunn-Thuan g, r, and i bands, which provide a good match with the plate and filter transmission curves (Weir et al. 1995a). Palomar Sky Survey plate detection limits are about $`23`$ magnitudes brighter than for the CCD images. For a typical low redshift cluster, galaxies with an apparent magnitude brighter than $`19`$ mag contain $`>90\%`$ of the total cluster luminosity, making photographic plates well suited for measuring total cluster light.
Several methods have been used for measuring total cluster masses, with generally consistent results. This suggests that the total cluster masses can be measured with reasonable accuracy, although systematic variations between the different methods do exist.
The oldest method of estimating cluster masses is based on the distribution of galaxy redshifts (e.g., the virial mass estimator). Assuming that the distribution of galaxies is similar to the distribution of the total mass, the cluster is in virial equilibrium and the velocity dispersions are isotropic, the virial mass of a cluster is related to the virial radius, $`r_v`$, and the line of sight projected velocity dispersion of galaxies, $`\sigma `$ by: $`M_v=3\sigma ^2r_v/G`$. This equation overestimates the total mass if the cluster is sampled to a radius smaller then $`r_v`$, since the surface pressure term in the virial theorem (2U+T=3PV) reduces the mass needed to bind the system. Moreover, if velocity anisotropies in the cluster exist, or the assumption that mass follows light does not hold, the virial mass estimator may produce misleading results (The & White 1986, Meritt 1987). For example, Bailey (1982) has shown that relaxing the mass-follows-light assumption can result in total cluster masses being considerably reduced. In such a case, a M/L<sub>V</sub> ratio as low as $`50h_{50}`$ is consistent with observed velocity dispersions in the Coma Cluster.
Using X-ray emission from clusters to measure the total masses has several advantages over virial mass estimators, since some of the assumptions involved can be observationally tested. Clusters have X-ray luminosities on the order of $`10^{43}10^{45}`$ ergs/sec, generated primarily by thermal bremsstrahlung from the hot intra-cluster gas that fills the deep gravitational potential wells (e.g., Jones & Forman 1984). Under the assumptions that this gas is supported by thermal pressure, is in hydrostatic equilibrium and spherically symmetric, the total cluster mass can be estimated from the gas density and temperature profiles (Bahcall & Sarazin 1977, Mathews 1978).
Previous optical and X-ray studies of groups and clusters (e.g., David et al. 1990, David et al. 1995) generally took into account the temperature structure for cool systems as measured by *ROSAT*. For the hottest, richest clusters, the gas was often assumed to be isothermal and was characterized by the emission weighted temperature. Crude temperature maps can now be obtained using the spatially resolved ASCA spectra, after applying corrections for the point spread function (PSF) of the ASCA mirrors. For the majority of clusters in our sample, we find the temperature declines with radius. Other studies (Markevitch et al. 1998, Nevalainen et al. 1999) have also observed declining temperature profiles. Compared to using the measured temperature profile, the isothermal assumption underestimates the total mass at small radii, and overestimates it at large radii. In addition, azimuthal variations in the gas temperatures have been observed in a number of clusters that are indicative of recent merger activity (Donnelly et al. 1998, Henriksen et al. 2000). In such clusters, the assumption of hydrostatic equilibrium can break down, and the applicability of X-ray mass estimates can be questioned.
Direct measurements of cluster masses can be obtained from gravitational lensing distortions of background galaxies. However, only a limited number of systems have been studied using this method (e.g., Smail et al. 1995). For cooling-flow clusters where the assumption of hydrostatic equilibrium is expected to hold, Allen (1998) found good agreement between X-ray mass estimates and results from strong and weak lensing.
In this paper we improve on the earlier measurements of cluster mass-to-light ratios, gas mass fractions, the limits on the baryon mass fractions and the constraints on $`\mathrm{\Omega }_0`$ by accounting for the intracluster gas temperature profiles, as well as using better quality optical data for measuring cluster luminosities. We study clusters that show symmetric temperature decline with radius, supporting the assumption of hydrostatic equilibrium and spherical symmetry. Our sample consists of 7 clusters (A262, A426, A478, A1795, A2052, A2063, A2199) and one group (MKW4s). They were selected as members of an X-ray flux limited sample of clusters that were observed with ASCA and the ROSAT PSPC, and are within the limits of the Second Palomar Sky Survey ($`\delta >3^{}`$). The details of our sample are tabulated in Table 1.
In Section 2 of this paper we discuss the X-ray data reduction and analysis. In Section 3 we discuss the optical data analysis. The main results and discussion of their implications are presented in Section 4. We assume $`\mathrm{H}_0=50h_{50}\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$ and $`\mathrm{q}_0=0.5`$. All errors are 90% confidence.
## 2. X-ray data reduction and analysis
Under the assumptions that the intra-cluster medium is spherically symmetric and in hydrostatic equilibrium supported solely by thermal pressure, the gas density, $`\rho _g`$, temperature, $`T`$, pressure, $`p_g`$ and mass, $`M`$, are related by:
$$\frac{dp_g}{dr}=\rho _g\frac{GM(<r)}{r^2}$$
(1)
$$p_g=\frac{\rho _gkT}{\mu m_p}$$
(2)
Here $`\mu `$ is the mean molecular weight of the gas (we assume $`\mu =0.6`$), and $`k`$ is Boltzmannโs constant. The mass within a radius $`r`$ is then:
$$M(<r)=\frac{kT(r)}{\mu m_pG}\left(\frac{d\mathrm{log}\rho _g(r)}{d\mathrm{log}r}+\frac{d\mathrm{log}T(r)}{d\mathrm{log}r}\right)r$$
(3)
Hence, the mass depends on both the gas density and temperature profiles. For isothermal gas, the observed surface brightness (which can be accurately obtained from ROSAT PSPC data), is directly related to the gas density. The surface brightness outside the cooling flow usually follows a $`\beta `$-profile (Cavaliere & Fusco-Femiano 1976) with a fixed background $`B`$:
$$I(r)=I_0\left[1+\left(\frac{r}{a}\right)^2\right]^{3\beta +\frac{1}{2}}+B$$
(4)
Here $`\beta =\mu m_p\sigma _r^2/kT_g`$ is the ratio of energy per unit mass in galaxies to the energy per unit mass in gas, and $`\sigma _r`$ is the velocity dispersion. The parameters $`a`$, $`\beta `$, and the background $`B`$ are obtained from a least-squares fit to the X-ray data, and the gas density profile is then given by
$$\rho _g(r)=\rho _0\left[1+\left(\frac{r}{a}\right)^2\right]^{\frac{3}{2}\beta }$$
(5)
The error introduced by assuming an isothermal gas in the density profile calculation is not significant, since the fraction of the bolometric luminosity emitted in the Snowden bands R5-R7 used in our analysis (0.7-2.0 keV) varies little with temperature for all clusters in our sample.
The central density $`\rho _0`$ can be found from the surface brightness profile as follows: from the known central surface brightness of the cluster (the $`\beta `$ profile extrapolated to the cluster center), the cluster redshift, gas temperature, abundance, absorbing hydrogen column density, effective area of the ROSAT mirrors and quantum efficiency of the PSPC in the 0.7-2.0 keV range, we can calculate the emission integral $`EI=n_pn_e๐V`$. For an isothermal $`\beta `$ model the emission integral is:
$$EI=\pi ^{3/2}\frac{n_e}{n_p}n_0^2a^3\frac{\mathrm{\Gamma }(3\beta 3/2)}{\mathrm{\Gamma }(3\beta )}$$
(6)
where $`n_0`$ is the central proton density, $`\mathrm{\Gamma }`$ is the gamma function, and for an assumed typical elemental abundance 0.3 Solar, $`n_e/n_p=1.17`$ and $`\rho =1.35m_pn_0`$. The total gas mass can then be found by integrating Equation 5 over the total volume. The effect of the cooling flow on the total gas mass measured at radii of 1 Mpc or greater from the cluster core is $`<10\%`$, since most of the gas mass resides in outer regions of clusters.
The most accurate temperature profiles for our sample clusters are available for A426 (presented here), and A2199 (Markevitch et al. 1999). In both cases, the temperature outside the cooling core can be well approximated by a polytrope, $`T\rho _g^{\gamma 1}`$ with $`\gamma 1.2`$.
For a gas distribution given by Equation 5, the total mass enclosed in a sphere of radius $`r=xa`$ is:
$`M(<r)`$ $`={\displaystyle \frac{kT(r)}{Gm_p\mu }}{\displaystyle \frac{3\beta \gamma r^3}{a^2+r^2}}=`$
$`=3.70\times 10^{13}M_{}{\displaystyle \frac{0.60}{\mu }}{\displaystyle \frac{T(r)}{1\mathrm{keV}}}{\displaystyle \frac{a}{1\mathrm{Mpc}}}{\displaystyle \frac{3\beta \gamma x^3}{1+x^2}}`$ (7)
Here $`T`$ is the real temperature, rather than a projection on the plane of the sky. Markevitch et al. (1999) has shown that as long as the temperature is proportional to a power of density, and the density follows a $`\beta `$-model, the real temperature differs from the projected temperature only by a constant factor, given by:
$$\frac{T_{\mathrm{proj}}}{T}=\frac{\mathrm{\Gamma }\left[\frac{3}{2}\beta (1+\gamma )\frac{1}{2}\right]\mathrm{\Gamma }(3\beta )}{\mathrm{\Gamma }\left[\frac{3}{2}\beta (1+\gamma )\right]\mathrm{\Gamma }(3\beta \frac{1}{2})}$$
(8)
This correction factor is in the range 0.9 to 0.98 for all clusters in our sample.
### 2.1. ROSAT data analysis
Archival ROSAT PSPC images were reduced using the standard analysis software (Snowden et al. 1994) that flat-fields the images and excludes periods of high particle background, as well as a period of 15 seconds after the high voltage is turned on. In order to maximize the signal-to-noise ratio, we use only Snowden energy bands R5-R7 corresponding to $`0.72.0`$ keV.
We fit the surface brightness profiles with $`\beta `$ models, with the core radius, $`\beta `$, the background, and the normalization as free parameters. Since we are primarily interested in the gas properties at large radii, the surface brightness profiles were fitted only outside twice the cooling radii taken from White, Jones & Forman (1997). An acceptable $`\chi ^2`$ cannot be obtained when the cooling flow region is included. Point sources were excluded from all images manually.
The results of the fitting procedure are shown in Table 2. Our determinations agree very well with earlier results from both ROSAT (Vikhlinin et al. 1999) and Einstein (Jones & Forman 1999) observations.
### 2.2. ASCA data analysis
The ASCA X-ray observatory (Tanaka, Inoue & Holt 1994) spatially resolved spectral data can be used to constrain the gas temperatures at different regions of the clusters. The ASCA mirrors have an energy and position dependent PSF that needs to be correctly taken into account. Two independent methods that correct for the PSF (Churazov et al. 1996; Markevitch et al. 1998) have been used in the past and were found to be in very good agreement (Donnelly et al. 1998). The first method approximates the ASCA PSF as having a core and broad wings. It uses the exact PSF correction for the core (inner 6โ), and a Monte Carlo correction for the scattered light in the wings of the PSF. The second method simultaneously fits temperatures in all selected regions, taking into account the observed surface brightness in each region and using the actual measured PSF.
We have adopted temperature profiles for three of the clusters (A478, A1795 and A2199) previously generated by Markevitch et al. (1998, 1999) using the second method described above. For the remaining five objects (A262, A426, A2052, A2063 and MKW4s), we have constructed temperature profiles using the first method. To check that the two methods for generating temperature profiles are consistent, we have constructed a temperature profile for A399 (Figure 1), which was presented by Markevitch et al. (1998). Applying the two methods yields results that agree well within their uncertainties. (The slightly different temperatures measured at large radii may be due to the small azimuthal asymmetry in the temperature structure present in this cluster.) A sample temperature profile (A426) we generated using the first method and a corresponding total mass profile obtained by fitting the temperature profile with a polytropic function are shown in Figure 2. A temperature profile for A426 has also been measured by Eyles et al. (1991) using an X-ray telescope flown on the *Spacelab 2* mission and is in excellent agreement with our measurement.
Gas mass fractions for all clusters in our sample are plotted as functions of enclosed mass and radius in Figure 3 and the results of our fitting are given in Table 3. The gas mass fractions reach $`0.150.25h_{50}^{3/2}`$ at a radius of 1 Mpc.
## 3. Optical data reduction and analysis
For measuring cluster luminosities, we use the Digitized Second Palomar Sky Survey (DPOSS), calibrated with photometric CCD images taken at the Palomar 60-in. telescope in the Gunn-Thuan g, r, and i bands (Weir et al. 1995a, Djorgovski et al. 1998).
### 3.1. Plate processing
The conversion of photographic plate emulsion density to intensity using the plate densitometry spots is described in Weir et al. (1995b).
The Sky Image Cataloging and Analysis Tool (SKICAT) has been developed to detect objects and perform star/galaxy classification on both DPOSS plates and CCD calibration data (Weir et al. 1995b). SKICAT is presently optimized for measuring fainter objects than $`m`$ 16 mag. Clusters in our sample contain galaxies brighter than this limit; hence we have used SExtractor (Bertin & Arnouts, 1995) for detecting objects and classifying stars and galaxies.
### 3.2. Photometric calibration
CCD images were obtained under photometric conditions for A262 (taken on 13 Dec 1998), A426 (12 Feb 1995), A478 (18 Sep 1998), A1795 (18 Jul 1999), A2063 (12 Jul 1999), and A2199 (18 Jul 1999) in the Gunn-Thuan g, r, i bands. To provide photometric calibration for A2052 and MKW4s, we use CCD images of different Abell clusters located on the respective plates near the clusters of interest: A2063 to calibrate A2052, A1495 (17 May 1998) to calibrate MKW4s. In order to correct the calibration of A2052 and MKW4s for vignetting effects, we median averaged $`100`$ POSS-II fields to obtain a vignetting map. The luminosity correction is on the order of a percent for both clusters. An example of a calibration transformation derived for A478 is shown in Figure 4.
In order to obtain the rest-frame galaxy luminosities, we need to correct for galactic absorption and k-dimming. Extinction corrections for clusters in our sample are given in Table 1 , taken from Schlegel et al. (1998).
K-corrections depend on spectral type, which can be related to galaxy morphological type. Since an automated morphological classification of galaxies in our sample is beyond the scope of this work, as well as very problematic at faint magnitudes, we assume a morphological composition and adopt k-corrections in a statistical manner (Table 4). A sample of 55 nearby rich clusters in the redshift range of our interest has been studied (Dressler 1980; Whitmore, Gilmore & Jones 1993, Dressler et al. 1997) and the morphological fractions determined as a function of the density of the environment. We follow Dressler et al. (1997) and adopt the following morphological fractions to be typical of the clusters in our sample: 25%:40%:35% for E:S0:Sp. The k-corrections are calculated using model galaxy SEDs from Small (1996) for different galaxy morphological types and the Gunn-Thuan g, r, i filter bandpasses (Weir et al. 1995a). We found the k-corrections from Small (1996) to be in agreement with an independent study by Fukugita et al. (1995).
It should be noted that for the majority of clusters in our sample, the k-correction in all three bands (g, r, and i) and all morphological types is no larger than 0.14 mag, and for about half of our sample, the k-correction is below 0.05 mag. Therefore, the exact morphological fraction is not critical, particularly in the r and i bands, where the differences in the k-corrections between the different morphological types are quite small at low redshifts. For example, had we assumed a spiral fraction of 20% rather than 35%, the statistically combined k-correction would change by no more than 0.02 mag for all clusters in our sample. Evolutionary effects also are insignificant due to the low redshift of our sample.
To convert apparent magnitudes to absolute magnitudes, we use the standard relation:
$$M=mDMEk$$
where $`DM`$ is distance modulus, $`E`$ is the galactic absorption and $`k`$ is the k-correction. Assuming $`q_0=0.5`$, the distance modulus, $`DM`$, is:
$$DM=43.89+5\mathrm{log}(z)5\mathrm{log}(h_{50})+0.54z$$
A deceleration parameter $`q_0=0`$ would increase the distance modulus $`DM`$ by $`0.03`$ mag at $`z=0.05`$ and by $`0.08`$ mag at $`z=0.15`$. A mean cluster redshift error of $`0.5\%`$ results in an absolute magnitude error of $`0.01`$, well below other random and systematic errors.
For comparison with other studies, we convert our Gunn g magnitudes to the Johnson $`V`$ band using the relation $`V=g0.030.42(gr)`$ (Windhorst et al. 1991). A mean $`gr`$ color for low redshift clusters is $`gr=0.3`$, giving $`V=r+0.14`$
### 3.3. Luminosity function determination
The values of galaxy cluster and group luminosities (e.g., Oemler 1974, Dressler 1978a,b; Bucknell et al. 1979; Lugger 1986; Oegerle et al. 1986; Ferguson & Sandage 1990) used in previous mass-to-light ratio studies date back to the first generation Palomar Sky Survey plates, or plates of similar grade taken elsewhere. In many studies only a small number of objects was used for photometric calibration, and star-galaxy classification was performed using simple two-parameter classifiers that are outperformed by more recent methods. In some studies object detection was performed by visual inspection.
Cluster luminosity functions have recently been studied using photometric CCD images (e.g., Lopez-Cruz et al. 1997). However, the number of clusters thus studied is still small, and at low redshifts the volume sampled is limited. Photographic plates still remain the optimal way of studying large samples of clusters over large areas of the sky.
Here we present the luminosity functions (LFs) for our sample of 8 relaxed clusters and one group, obtained from the digitized second generation Palomar Sky Survey plates, calibrated with CCD images in the Gunn-Thuan g, r, i system, and sampling 1 Mpc from the cluster centers.
#### 3.3.1 Background subtraction
Different LF studies have taken different paths in estimating background counts. Some have used values obtained in other independent studies, where different filters, angular coverage, or a different definition for galaxy magnitudes were used. These factors introduce errors that were estimated to contribute to a total uncertainty of $`\pm 50\%`$ in the background correction (Oemler 1974, Lugger 1986, Colless 1989). Lower values in the background uncertainty were reported by Dressler (1978a) using Shane and Wirtanen counts (25% variation). Lopez-Cruz (1995) found a similar variation in $`R`$-band counts on scales $`0.4^{}`$
We have analyzed the background galaxy counts on POSS-II plates 725, and found a 19% variation on scales of $`0.5^{}`$ with a limiting *J* magnitude $`19.5`$. This is in agreement with the findings of Dressler (1978a) and the expected variation in the angular covariance function (Groth & Peebles 1975) on scales of 0.5. We thus assume the error in the background counts $`N`$ is the maximum of $`\sqrt{N}`$ and $`N/5`$.
Assuming Poisson uncertainties in the uncorrected galaxy counts, the error in the corrected counts is given by
$$\sqrt{N+\mathrm{max}(\sqrt{\text{N}};\text{N}/5)^2}.$$
The background subtracted differential LFs were fitted with the commonly used Schechter function (Schechter 1976):
$$n(L)dL=N^{}(L/L^{})^\alpha \mathrm{exp}(L/L^{})d(L/L^{})$$
Once the parameters $`L^{}`$, $`N^{}`$ and $`\alpha `$ have been determined, the total cluster luminosity is given by:
$$L_{\mathrm{clus}}=_0^{\mathrm{}}Ln(L)๐L=N^{}\mathrm{\Gamma }(\alpha +2)L^{}.$$
For the three lowest redshift clusters (A262, A426, A2199) where the absolute magnitude range sampled is the greatest, we found a sum of two Schechter functions greatly improves the chi-square of the fit. In such cases the slope of the brighter component was fixed at $`\alpha =1`$, the remaining parameters were left free.
Since we measure cluster masses within 1 Mpc from the cluster centers, to obtain the corresponding luminosities over these cluster volumes, we must correct for outlying cluster galaxies projected near the cluster center. To calculate this correction we need to know the galaxy number density and average galaxy luminosity as a function of distance from the cluster center. We fitted the galaxy number density profiles with $`\beta `$ models and found the coefficients to have a larger uncertainty, but to be consistent with the gas density fitting results. We thus assume that the distribution of galaxies follows the distribution of intracluster gas, and that the average galaxy luminosity is independent of the density of the environment. The correction factor is in the range $`0.900.97`$ and has the effect of decreasing the true total cluster luminosities. Some studies have suggested both of the underlying assumptions may be violated. However, the error this may introduce can be only of the order of a percent, since we sample to a radial distance about 5 times the typical cluster core radius.
In our determinations of cluster luminosities we assume the Schechter function is a universal LF valid over a large magnitude range from giant galaxies to dwarfs. We study the LFs of galaxies with apparent r magnitude brighter than $`19`$ mag (corresponding to an absolute magnitude $`M_r17.5`$ to $`19.5`$ depending on the cluster redshift). The giant galaxies contribute most of the cluster luminosity. Typically, galaxies with $`M_r<19`$ comprise 80-90% of the total cluster luminosity, with the exact number depending on $`M^{}`$ and $`\alpha `$.
Some studies have found the Schechter function does not describe the cluster LF well at the faint end. Trentham (1998) studied B-band LFs of 9 Abell clusters and showed that LFs tend to flatten for $`18<M_B<16`$ and then rise for fainter galaxies, with slopes varying in the range $`1.3<\alpha <1.8`$. However, for our purposes this effect is negligible, since the dwarf galaxies contribute only a small fraction of the total light. Assuming the LF is described by a Schechter function with $`M^{}22`$ and $`1.4<\alpha <1.0`$ for $`M_r<17`$ (as suggested by Trentham 1998), the effect of a faint end slope varying in the range $`1>\alpha >2`$ results in a negligible change in the total cluster luminosity ($`1\%`$). The results of the fitting procedure are shown in Table 5.
## 4. Mass-to-light ratios and constraints on $`\mathrm{\Omega }`$
The mass to light ratio, $`\mathrm{M}/\mathrm{L}`$, is used to parameterize the amount of dark matter on various scales. $`\mathrm{M}/\mathrm{L}`$ increases with scale from galaxies to groups and clusters (Bahcall, Lubin, & Dorman 1995). However, a flattening of the $`\mathrm{M}/\mathrm{L}`$ vs. scale relation has been observed on scales beyond clusters, as discussed in the introduction. Assuming that the mass-to-light ratios of clusters are representative of the whole Universe, the mass density of the Universe can be calculated from the observed mean luminosity density of the Universe and $`\mathrm{M}/\mathrm{L}`$ of clusters.
The median $`\mathrm{M}/\mathrm{L}`$ of our sample is $`\mathrm{M}/\mathrm{L}_\mathrm{V}100h_{50}\mathrm{M}_{}/\mathrm{L}_{}`$ (Table 5). The mean luminosity density of the universe is $`1\times 10^8h_{50}\mathrm{L}_{}\mathrm{Mpc}^3`$ (Efstathiou et al. 1988). This gives a universal mass density of $`\rho _m7\times 10^{31}h_{50}^2\mathrm{g}\mathrm{cm}^3`$. With a critical density of $`\rho _{\mathrm{crit}}5\times 10^{30}h_{50}^2\mathrm{g}\mathrm{cm}^3`$, we obtain $`\mathrm{\Omega }_00.15`$.
Our mass-to-light ratios within 1 Mpc are slightly lower than previous results that used X-ray mass estimates: $`\mathrm{M}/\mathrm{L}_\mathrm{V}100h_{50}`$ solar units compared to $`120150h_{50}`$ (Cowie 1987, David et al. 1995). This discrepancy may be due in part to the inability in previous work to correct for temperature structure. With the advent of spatially resolved spectral measurements from ASCA we would expect a difference in the results especially if temperature gradients are common. Second, we have used larger datasets for calibrating the plate magnitudes, in comparison with earlier studies. As a result, we expect the combined photometric properties of larger samples of galaxies (such as the total luminosity) to be more accurate estimates of the true values. There is only one cluster (A262) studied both in this paper, and by David et al. (1995). The mass-to-light ratios measured are in good agreement.
$`\mathrm{M}/\mathrm{L}_\mathrm{V}`$ for most clusters in our sample are also lower than estimates based on the virial mass estimator, which typically yield $`\mathrm{M}/\mathrm{L}_\mathrm{V}125180h_{50}\mathrm{M}_{}/\mathrm{L}_{}`$ (e.g., Girardi et al. 1999, Carlberg et al. 1996). As argued in the introduction, virial mass estimates can be misleading if substructure is present, the assumption that mass follows light fails, or when the volume sampled does not extend to the virial radius, which is the case in many studies.
Our analysis shows that $`\mathrm{M}/\mathrm{L}`$ is roughly independent of cluster mass as characterized by richness or temperature (Figure 5 and Table 1.) This is contrary to the popular belief that mass-to-light ratios increase with richness from groups to clusters, and is in agreement with the findings of David et al. (1995).
Standard Big Bang nucleosynthesis limits the baryon density of the universe to $`\mathrm{\Omega }_b=0.076\pm 0.004h_{50}^2`$, where $`\mathrm{\Omega }_b=f_b\mathrm{\Omega }_0`$, $`f_b`$ is the baryon mass fraction (Walker et al. 1991, White et al. 1993, Tytler et al. 1996, Kirkman et al. 2000). In Section 2 we showed the gas mass fraction reaches $`0.150.25h_{50}^{3/2}`$ at a 1 Mpc radius and tends to increase further towards larger radii, with stars contributing only a few percent of the baryon mass throughout. If we assume the standard Big Bang nucleosynthesis calculations correctly predict the expected baryon fraction and that the gas fraction found in clusters of galaxies is representative of the baryon fraction in the Universe, as White et al. (1993) and David et al. (1995) have done, we also can place an upper limit on $`\mathrm{\Omega }_0`$: $`\mathrm{\Omega }_0<0.076f_b^1h_{50}^{1/2}`$. For our best estimate $`f_b=0.25`$ (taking the upper limit to account for gas fractions increasing beyond the region surveyed), we obtain $`\mathrm{\Omega }_0<0.30h_{50}^{1/2}`$, which is consistent with the constraint on $`\mathrm{\Omega }_0`$ from mass-to-light ratios for a presently favored value of $`H_0=65\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$. We note that a larger $`\mathrm{\Omega }_0`$ is allowed if a lower gas mass fraction is adopted. For our lower limit $`f_b=0.15`$, we obtain $`\mathrm{\Omega }_0<0.51h_{50}^{1/2}`$.
## 5. Conclusion
We have investigated several fundamental properties of a sample of 7 Abell clusters and one group. We have utilized the Digitized Second Palomar Sky Survey optical data and photometric CCD images for constraining cluster luminosities, along with ROSAT X-ray data and ASCA spectra for constraining total and gas masses.
We have measured the median cluster mass-to-light ratios within 1 Mpc to be $`\mathrm{M}/\mathrm{L}_\mathrm{V}100h_{50}\mathrm{M}_{}/\mathrm{L}_{}`$, corresponding to $`\mathrm{\Omega }_00.15`$. This is slightly lower than found in other studies that used X-ray mass estimates, and lower compared to results based on virial mass estimates.
We have measured the gas mass fractions in the range 0.1-1 Mpc, and found these to approach $`0.150.25h_{50}^{3/2}`$ towards the cluster virial radii. Using the standard Big Bang nucleosynthesis calculations, assuming that the baryon fraction seen in clusters to within the virial radius of clusters is representative of the overall baryon fraction in the Universe, we find the total matter density of the universe to be $`\mathrm{\Omega }_0<0.30h_{50}^{1/2}`$.
Our two determinations of $`\mathrm{\Omega }_0`$ are in agreement. Our results also are consistent within their uncertainties with other independent measurements of $`\mathrm{\Omega }_0`$, such as the evolution of cluster abundance as a function of redshift (Bahcall 1999), microwave background fluctuations based on the COBE satellite results assuming both OCDM and LCDM models (e.g., Cayรณn et al. 1996), or measurements using distant supernovae (e.g., Perlmutter et al. 1999, Riess et al. 1998)
As we enter the era of large format optical CCDs, it will become possible to study the luminosity functions of large samples of clusters out to the virial radii and reaching fainter magnitudes than photographic plates. New X-ray missions with better spectral and spatial resolution, such as Chandra and XMM, will better constrain the properties of the ICM in clusters, which will decrease the present uncertainties on the mass-to-light ratios and limits on the baryon fraction.
V. H. was partially supported by the Caltech SURF fellowship. V. H. would like to thank the Harvard-Smithsonian center for Astrophysics for hospitality. C. J. and R.H.D. acknowledge support form the Smithsonian Institute and NASA contract NAS8-39073. The DPOSS cataloging effort is supported by a generous grant from the Norris foundation. V. H. would further like to thank A. Vikhlinin and M. Markevitch for useful comments.
|
warning/0006/math0006099.html
|
ar5iv
|
text
|
# Equivariant resolution of points of indeterminacy
## 1. Introduction
Throughout this note we shall work over an algebraically closed field of characteristic 0. All algebraic varieties, schemes, groups, and all maps between them will be defined over $`k`$. The main objects of interest for us will be algebraic varieties with a $`G`$-action; we will refer to them as $`G`$-varieties. A $`G`$-equivariant morphism between two such varieties will be called a morphism of $`G`$-varieties. The terms โrational map of $`G`$-varietiesโ, โbirational morphism of $`G`$-varietiesโ, โbirational isomorphism of $`G`$-varietiesโ, etc., are defined in a similar manner.
Hironakaโs theorem on elimination of points of indeterminacy (see \[Hi, ยง0.5, Question E and Main Theorem II\]) asserts that every rational map $`f:XY`$ can be resolved into a regular map by a sequence of blowups $`\pi :X_m\mathrm{}X_0=X`$ with smooth centers. In other words, $`\pi `$ can be chosen so that the composition $`f\pi `$ is regular. The purpose of this paper is to prove the following equivariant version of this result.
###### Theorem 1.
Let $`f:XY`$ be a rational map of $`G`$-varieties where $`Y`$ is complete. Then there is a sequence of blowups
(1)
$$\pi :X_mX_{m1}\mathrm{}X_1X_0=X$$
with smooth $`G`$-invariant centers such that the composition $`f\pi `$ is regular.
Our proof will rely on canonical resolution of singularities. Along the way we prove an equivariant form of Chowโs lemma (Proposition 5), generalizing a theorem of Sumihiro (\[Su, Theorem 2\]).
The second author warmly thanks Institute of Mathematics of Hebrew University for its hospitality during 1999/2000.
## 2. Birational morphisms as blowups
The following result is an equivariant analogue of \[Ha, Theorem 7.17\].
###### Proposition 2.
Let $`f:X^{}X`$ be a birational proper morphism of $`G`$-varieties, where $`X`$ is smooth and $`X^{}`$ is quasiprojective. Then there exists a $`G`$-invariant sheaf of ideals $``$ on $`X`$ such that $`X^{}`$ is the blowup of $``$.
###### Proof.
Let $`\sigma :G\times XX`$ be the given action of $`G`$ on $`X`$ and $`\mathrm{pr}_2:G\times XX`$ be the projection onto the second factor.
By a theorem of Kambayashi \[Ka\], there is an action of $`G`$ on the projective space $`^n`$ (via a representation $`G\mathrm{PGL}_{n+1}`$) and a $`G`$-equivariant embedding $`X^{}^n`$; this yields a $`G`$-equivariant embedding $`i:X^{}^n\times X`$.
Here $`^n\times X`$ is a projective space over $`X`$; set $`=i^{}๐ช_{^n\times X}(1)`$ and $`๐ฏ=_{d=0}^{\mathrm{}}f_{}(^d)`$, where $`^0=๐ช_{^n\times X}`$. Let $`๐ฏ_1=f_{}`$ be the component of $`๐ฏ`$ of degree one. The action of $`G`$ on $`^n\times X`$ yields a $`G`$-linearization of the sheaf $`๐ฏ_1`$, i.e., an isomorphism $`\sigma ^{}๐ฏ_1\stackrel{}{=}\mathrm{pr}_2^{}๐ฏ_1`$ which satisfies the same cocycle condition as in the definition of $`G`$-linearization of an invertible sheaf (see, e.g., \[MFK, Definition 1.6\]); informally speaking, $`G`$ acts on the pair $`(X,๐ฏ_1)`$.
We refer to the proof of \[Ha, Theorem 7.17\] for the following facts:
1. After replacing the embedding $`i`$ by its $`e`$-fold embedding for some positive integer $`e`$ (thus replacing $``$ by $`^e`$), we may assume that the graded $`๐ช_X`$-algebra $`๐ฏ`$ is generated by $`๐ฏ_1`$.
2. $`X^{}\stackrel{}{=}\mathrm{Proj}๐ฏ`$.
3. Assume $`๐ฏ`$ is generated by $`๐ฏ_1`$ as in (1). If there is an invertible sheaf $``$ on $`X`$ and a sheaf of ideals $``$ on $`X`$ such that $`\stackrel{}{=}๐ฏ_1`$, then $`X^{}`$ is isomorphic to the blowup of $``$.
The variety $`X`$ is smooth, and hence, for any sheaf of ideals $``$ on $`X`$ of rank one without torsion, its dual $`^{}=\mathrm{om}(,๐ช_X)`$ is an invertible sheaf. (To see this, note that locally at any point $`xX`$, the generator of $`_x^{}`$ is given by the homomorphism $`_x๐ช_{X,x}`$ which maps the generators of $`_x`$ as an $`๐ช_{X,x}`$-module of rank one, into elements of $`๐ช_{X,x}`$ not having a nontrivial common multiple; such homomorphism is unique up to an invertible multiple since the local ring $`๐ช_{X,x}`$ is regular, and hence, factorial.)
Thus the second dual $`๐ฏ_1^{}`$ is an invertible sheaf, and we have an embedding $`๐ฏ_1๐ฏ_1^{}`$. The $`G`$-linearization of $`๐ฏ_1`$ yields a $`G`$-linearization of $`๐ฏ_1^{}`$, and the above embedding is, in fact, an embedding of $`G`$-linearized sheaves. Taking $`=(๐ฏ_1^{})^1`$, we see that $`X^{}`$ is isomorphic to the blowup of the sheaf of ideals $`=๐ฏ_1(๐ฏ_1^{})^1`$.
The $`G`$-linearizations of $`๐ฏ_1`$ and $`๐ฏ_1^{}`$ yield a $`G`$-linearization of $``$ and a $`G`$-linearized embedding $`=๐ฏ_1(๐ฏ_1^{})^1๐ฏ_1^{}(๐ฏ_1^{})^1=๐ช_X`$. This shows that $``$ is a $`G`$-invariant sheaf of ideals on $`X`$. โ
## 3. Canonical simplification of a finite collection of ideals
One of the main resolution theorems of Hironaka \[Hi, Main Theorem II\] asserts that any finite collection of sheaves of ideals $`\{_i\}`$ can be โsimplifiedโ by a finite sequence $`\pi :X_m\mathrm{}X_0=X`$ of blowups with smooth centers. In other words, the sequence of blowups can be chosen so that $`\pi ^{}_i`$ is locally principal for each $`i`$, and is locally generated by a monomial with respect to a normal crossing divisor.
Bierstone and Milman \[BM, Theorem 1.10\] proved that any sheaf of ideals $``$ on an algebraic variety $`X`$ can be โsimplifiedโ in this sense in a canonical way, so that the sequence $`\pi `$ is canonically defined; in particular, every automorphism of $`X`$ preserving $``$ lifts to the entire sequence; see \[BM, Remark 1.5\]. This immediately implies that any finite ordered collection of sheaves of ideals $`\{_i\}`$ can be โsimplifiedโ in a canonical way.
In this section we will show that a finite unordered collection of sheaves of ideals can be simplified in a similar manner. This result will be used in the proof of Theorem 1.
###### Proposition 3.
Let $`X`$ be an algebraic variety and $`\{_1,\mathrm{},_n\}`$ be a finite collection of sheaves of ideals on $`X`$. Then there exists a sequence of blowups
(2)
$$\pi :X_mX_{m1}\mathrm{}X_1X_0=X$$
such that $`\pi ^{}_i`$ is locally principal for each $`i`$ and any automorphism of $`X`$ that preserves (but possibly non-trivially permutes) the collection $`\{_1,\mathrm{},_n\}`$ lifts to the entire sequence (2).
###### Proof.
To motivate our construction, we begin with the following observation. Let $`V(_i)`$ be the subscheme of $`X`$ cut out by $`_i`$. If the subschemes $`V(_i)`$ are pairwise disjoint, i.e., if $`_i+_j=๐ช_x`$ for any $`ij`$, then the proposition follows immediately from \[BM, Theorem 1.10 together with Remark 1.5\]: indeed, a sequence of blowups that simplifies their intersection $`_1\mathrm{}_n`$, will simplify each $`_i`$.
The idea of the proof is to reduce the general case to the case where every $`i`$-fold intersection of $`V(_1),\mathrm{},V(_n)`$ is empty (i.e. the sum of any $`i`$ of the sheaves $`_1,\mathrm{},_n`$ equals $`๐ช_X`$) first for $`i=n`$, then for $`i=n1`$, etc., until we reach $`i=2`$. We use descending induction on $`i`$. For the base case we can take $`i=n+1`$, where the condition we are interested in is trivially satisfied.
Let $`S_\mathrm{\Lambda }=_{j\mathrm{\Lambda }}_j`$, where $`\mathrm{\Lambda }`$ is a subset of $`\{1,\mathrm{},n\}`$. For the induction step, assume
(3) $`S_\mathrm{\Lambda }=๐ช_X`$ whenever $`|\mathrm{\Lambda }|=i`$,
for some $`i2`$. Set $`๐ฅ=_{|\mathrm{\Omega }|=i1}S_\mathrm{\Omega }`$. Note that by our assumption $`S_{\mathrm{\Omega }_1}+S_{\mathrm{\Omega }_2}=๐ช_X`$ for any two distinct subsets $`\mathrm{\Omega }_1`$ and $`\mathrm{\Omega }_2`$ of $`\{1,\mathrm{},n\}`$ of cardinality $`i1`$, so that $`V(๐ฅ)`$ is the disjoint union of $`V(S_\mathrm{\Omega })=_{j\mathrm{\Omega }}V(_j)`$ with $`|\mathrm{\Omega }|=i1`$.
Let $`\pi :X^{}X`$ be the canonical simplification of the sheaf $`๐ฅ`$; as we have seen above, $`\pi `$ simplifies each $`S_\mathrm{\Omega }`$ with $`|\mathrm{\Omega }|=i1`$. For $`j=1,\mathrm{},n`$, denote the conductor $`(\pi ^{}_j):(\pi ^{}๐ฅ)`$ by $`_j^{}`$; it is natural to think of $`_j^{}`$ as a โweak transformโ of $`_j`$. The stalk of this sheaf of ideals at a (not necessarily closed) point $`xX^{}`$ is described as follows.
If $`xV(\pi ^{}_j)`$ then $`(_j^{})_x=(\pi ^{}_j)_x=๐ช_{x,X^{}}`$. If $`xV(\pi ^{}_j)`$ and $`xV(\pi ^{}S_\mathrm{\Omega })`$ for some $`\mathrm{\Omega }`$ satisfying $`|\mathrm{\Omega }|=i1`$, then such $`\mathrm{\Omega }`$ is unique and $`j\mathrm{\Omega }`$ (otherwise $`V(\pi ^{}S_{\mathrm{\Omega }\{j\}})`$ would be nonempty, contrary to (3)). Thus in this case $`\pi ^{}_j\pi ^{}S_\mathrm{\Omega }`$ and $`(_j^{})_x=(\pi ^{}S_\mathrm{\Omega })_x^1(\pi ^{}_j)_x`$, where $`(\pi ^{}S_\mathrm{\Omega })_x๐ช_{x,X^{}}`$ is a principal ideal. Finally, if $`xV(\pi ^{}_j)`$ and $`xV(\pi ^{}S_\mathrm{\Omega })`$ for any $`\mathrm{\Omega }`$ satisfying $`|\mathrm{\Omega }|=i1`$, then $`xV(\pi ^{}๐ฅ)`$ and $`(_j^{})_x=(\pi ^{}_j)_x`$. To summarize:
(4)
$$(_j^{})_x=\{\begin{array}{cc}(\pi ^{}S_\mathrm{\Omega })_x^1(\pi ^{}_j)_x\hfill & \text{ if }xV(\pi ^{}S_\mathrm{\Omega })\text{ for some }\mathrm{\Omega }\{1,\mathrm{},n\}\hfill \\ & \text{ such that }j\mathrm{\Omega }\text{ and }|\mathrm{\Omega }|=i1\text{,}\hfill \\ (\pi ^{}_j)_x\hfill & \text{ otherwise.}\hfill \end{array}$$
Consequently, for any sequence of blowups $`\pi ^{}:X^{\prime \prime }X^{}`$, the ideal $`(\pi ^{})^{}_j^{}`$ is locally principal if and only if the ideal $`(\pi ^{})^{}(\pi ^{}_j)`$ is locally principal. This reduces the problem of simplifying the collection $`\{_1,\mathrm{},_n\}`$ of sheaves of ideals on $`X`$ to the problem of simplifying the collection $`\{_1^{},\mathrm{},_n^{}\}`$ of sheaves of ideals on $`X^{}`$.
For $`\mathrm{\Lambda }\{1,\mathrm{},n\}`$, set $`S_\mathrm{\Lambda }^{}=_{j\mathrm{\Lambda }}_j^{}`$. We claim that
(5) $`S_\mathrm{\Lambda }^{}=๐ช_X^{}`$ whenever $`|\mathrm{\Lambda }|=i1`$.
We will prove this equality by showing that $`(S_\mathrm{\Lambda }^{})_x=๐ช_{X^{},x}`$ for every $`xX^{}`$. Indeed, if $`xV(\pi ^{}(_j))`$ for some $`j\mathrm{\Lambda }`$ then
$$๐ช_{X^{},x}=(\pi ^{}_j)_x(_j^{})_x(S_\mathrm{\Lambda }^{})_x,$$
as desired. On the other hand, if $`xV(\pi ^{}_j)`$ for every $`j\mathrm{\Lambda }`$, i.e., $`xV(\pi ^{}S_\mathrm{\Lambda })`$, then (4) tells us that
$$(S_\mathrm{\Lambda }^{})_x=\underset{j\mathrm{\Lambda }}{}(_j^{})_x=\underset{j\mathrm{\Lambda }}{}(\pi ^{}S_\mathrm{\Lambda })_x^1(\pi ^{}_j)_x=(\pi ^{}S_\mathrm{\Lambda })_x^1(\pi ^{}S_\mathrm{\Lambda })_x=๐ช_{X^{},x}.$$
We have thus reduced the problem of simplifying the collection $`\{_1,\mathrm{},_n\}`$ of sheaves of ideals on $`X`$, satisfying condition (3), to the problem of simplifying the collection $`\{_1^{},\mathrm{},_n^{}\}`$ of sheaves of ideals on $`X^{}`$, satisfying condition (5). This completes the induction step.
To finish the proof of the proposition, note that the sequence (2) of blowups constructed by the recursive algorithm we just described, depends on $`X`$ and the unordered collection $`\{_i\}`$ in a canonical way; see \[BM, Remark 1.5\]. Hence, any automorphism of $`X`$ that preserves the unordered collection $`\{_i\}`$, lifts to to the entire sequence (2), as claimed. โ
###### Remark 4.
Our proof also shows that each $`\pi ^{}_i`$ is generated by a monomial with respect to a normal crossing divisor. In other words, $`\pi `$ simplifies each $`_i`$ in the sense of Hironakaโs original definition; for details see \[BM, Remark 1.8\]. This assertion will not be used in the sequel; for this reason we did not include it in the statement of Proposition 3.
## 4. Equivariant Chow Lemma
In this section we will prove the following generalization of Chowโs lemma.
###### Proposition 5.
For every $`G`$-variety $`X`$, there exists a quasiprojective $`G`$-variety $`Z`$ and a proper birational morphism $`ZX`$. If $`X`$ is complete then $`Z`$ is projective.
Note that if $`G`$ is assumed to be connected, this result is a well-known theorem of Sumihiro \[Su, Theorem 2\]; see also \[PV, Theorem 1.3\]. The argument below reduces the general case to the case where $`G`$ is connected.
###### Proof.
The second assertion is an immediate consequence of the first: if $`X`$ is complete, $`ZX`$ is proper and $`Z`$ is quasiprojective then $`Z`$ is also complete and, hence, projective.
To prove the first assertion let $`G_0`$ be the connected component of $`G`$. Applying Sumihiroโs theorem to $`X`$, viewed as a $`G_0`$-variety, yields a quasiprojective $`G_0`$-variety $`Y`$ and a proper $`G_0`$-equivariant birational morphism $`f:YX`$.
For the rest of this proof we shall use set-theoretic notation: by a โpointโ we will always mean a closed point.
Recall that the homogeneous fiber product $`G_{G_0}Y`$ is the $`G`$-variety defined as the geometric quotient $`(G\times Y)/G_0`$ for the action of $`G_0`$ given by $`g_0(g,y)=(gg_0^1,g_0y)`$, where $`gG`$, $`g_0G_0`$ and $`yY`$; see \[PV, Section 4.8\]. We shall write $`[g,y]`$ for the element of $`G_{G_0}Y`$ represented by $`(g,y)G\times Y`$. Since $`G_0`$ has finite index in $`G`$, $`G_{G_0}Y`$ admits a more concrete description as a disjoint union of $`|G/G_0|`$ copies of $`Y`$. More precisely, if we choose a representative $`a_h`$ for each $`hG/G_0`$, we can explicitly identify $`G/G_0\times Y`$ and $`G_{G_0}Y`$ as abstract varieties, via $`(h,y)[a_h,y]`$. Moreover, if we define a $`G`$-action on $`G/G_0\times Y`$ by $`g(h,y)(\overline{g}h,(a_h^1g^1a_{\overline{g}h})y)`$ then $`(h,y)[a_h,y]`$ identifies $`G/G_0\times Y`$ and $`G_{G_0}Y`$ as $`G`$-varieties. Here $`\overline{g}`$ is the image of $`g`$ in $`G/G_0`$ and $`(a_h^1g^1a_{\overline{g}h})y`$ is well-defined because $`a_h^1g^1a_{\overline{g}h}`$ is an element of $`G_0`$.
Let $`\alpha :G_{G_0}YX`$ and $`\beta :G_{G_0}YG/G_0`$ be the maps of $`G`$-varieties given by $`\alpha :[g,y]gf(y)`$ and $`\beta :[g,y]\overline{g}`$. (Here $`G`$ acts on $`G/G_0`$ by left multiplication.) These maps are shown in the diagram below.
$$\begin{array}{ccccc}& & G_{G_0}YG/G_0\times Y& & \\ & \alpha & & \beta s& \\ X& & & & G/G_0\end{array}$$
Let $`S`$ be the set of all sections $`s`$ of $`\beta `$. Note that if we identify $`G_{G_0}Y`$ with $`G/G_0\times Y`$ as above, then $`\beta :G/G_0\times YG/G_0`$ is the projection to the first factor. Thus $`SY^{|G/G_0|}`$ as an abstract variety. Moreover, since $`\beta `$ is $`G`$-equivariant, $`G`$ acts on this variety by $`g:st`$, where $`s,tS`$ and $`t(h)=gs(\overline{g}^1h)`$ for any $`hG/G_0`$.
Let $`Z`$ be the closed $`G`$-invariant subvariety of $`S`$ consisting of those sections $`s:G/G_0G_{G_0}YG/G_0\times Y`$ with the property that $`\alpha s(G/G_0)`$ is a single point of $`X`$; we shall denote this point by $`x_s`$.
We claim that the morphism $`\varphi :ZX`$ given by $`sx_s`$ has the properties asserted in the proposition. Indeed, since $`Y`$ is quasiprojective, and $`Z`$ is a closed subvariety of $`SY^{|G/G_0|}`$, $`Z`$ is quasiprojective as well.
To show that $`\varphi `$ is a birational morphism, assume the birational morphism $`f:YX`$ is an isomorphism over a dense open subset $`UX`$. Then $`V=_{hG/G_0}a_hU`$ is also a dense open subset of $`X`$, and for every $`xV`$, $`\alpha ^1(x)=\{[a_h,f^1(a_{h^1}x)]:hG/G_0\}`$; it is the image of the unique section $`s_xS`$ satisfying $`x_{s_x}=x`$. This section is given by $`s_x(h)=[a_h,f^1(a_h^1x)]`$, and the morphism $`VZ`$, $`xs_x`$, is a two-sided rational inverse to $`\varphi `$. โ
## 5. Proof of Theorem 1
We begin with two reductions. First of all, we may assume without loss of generality that $`Y`$ is projective. Indeed, Proposition 5 yields a projective $`G`$-variety $`Y^{}`$ and a $`G`$-equivariant birational morphism $`u:Y^{}Y`$. We can now replace $`Y`$ by $`Y^{}`$ and $`f`$ by $`f^{}=u^1f:XY^{}`$ If we can construct a sequence of blowups $`\pi `$, as in (1), so that $`f^{}\pi `$ is regular, then $`f\pi `$ is regular as well, i.e., the same sequence of blow ups will resolve the indeterminacy locus of $`f`$.
Secondly, we may assume that $`X`$ is smooth. Indeed, let
(6)
$$X_l\mathrm{@}>\pi _l>>\mathrm{}\mathrm{@}>\pi _1>>X_0=X,$$
be the canonical resolution of singularities of $`X`$, as in \[V, Theorem 7.6.1\] or \[BM, Theorem 13.2\]. Here $`X_l`$ is smooth, the centers $`C_iX_i`$ are smooth and $`G`$-invariant, and the action of $`G`$ lifts to the entire resolution sequence (6). Replacing $`X`$ by $`X_l`$, we may assume that $`X`$ is smooth.
From now on we will assume $`X`$ is smooth and $`Y`$ is projective. Let $`G_0`$ be the connected component of $`G`$. As $`X`$ is smooth, it is normal, and we can apply a theorem of Sumihiro (see \[Su, Theorem 1\] or \[KKLV, Theorem 1.1\] or \[PV, Theorem 1.2\]) which yields a finite covering $`X=U_1\mathrm{}U_d`$, where each $`U_i`$ is a $`G_0`$-invariant quasiprojective open subvariety of $`X`$.
For each $`i`$, let $`Z_i`$ be the closure of the graph of $`f|_{U_i}:U_iY`$ in $`U_i\times Y`$; it is a quasiprojective variety, and the projection $`h_i:Z_iU_i`$ is a proper birational morphism of $`G`$-varieties. By Proposition 2, we can find a $`G_0`$-invariant sheaf of ideals $`_i`$ on $`U_i`$ such that $`Z_i`$ is isomorphic to the blowup of $`_i`$. Let $`_i^{}`$ be the maximal sheaf of ideals on $`X`$ such that $`_i^{}|_{U_i}=_i`$; as $`_i^{}`$ is unique, it is $`G_0`$-invariant. The $`G`$-invariant collection of sheaves $`\{g^{}_i^{}|gG,i=1,\mathrm{},d\}`$ on $`X`$ is finite (it contains no more than $`|G/G_0|`$ sheaves for each $`i`$). Proposition 3 yields a $`G`$-equivariant sequence of blowups
$$\pi :X_mX_{m1}\mathrm{}X_1X_0=X$$
with the property that the pullback $`\pi ^{}_i^{}`$ is locally principal for each $`i`$; hence, the composition $`h_i^1\pi :X_mXZ_i`$ is regular on $`\pi ^1(U_i)`$. This implies that $`f\pi :X_mXZ_iY`$ is regular on $`\pi ^1(U_i)`$ for each $`i`$, and hence, on all of $`X_m`$. โ
###### Remark 6.
Note that if $`Y`$ is assumed to be projective in the statement of Theorem 1 then Proposition 5 is not needed in the proof. On the other hand, if $`G`$ is assumed to be connected, then Proposition 3 may be replaced by \[BM, Theorem 1.10 together with Remark 1.5\] and Proposition 5 may be replaced by \[Su, Theorem 2\].
|
warning/0006/hep-ph0006296.html
|
ar5iv
|
text
|
# A possible hadronic excess in ๐โข(2โข๐) decay and the ๐โข๐ puzzle
## I Introduction
Absence of the $`\rho \pi `$ decay mode of $`\psi (2S)`$ has defied a theoretical explanation for more than a decade. The recent measurement by BES Collaboration has confirmed the absence of $`\rho \pi `$ with even a higher precision, setting its upper bound at a level of factor more than 60 below what one naively expects from the decay $`J/\psi \rho \pi `$. The measurement of other decay modes by BES seems to rule out all possible resolutions for the $`\rho \pi `$ puzzle that have so far been proposed by theorists. For instance, the large $`\omega \pi `$ branching contradicts the helicity suppression with or without large intrinsic charm. A vector glueball near the $`J/\psi `$ mass, if it should exist, can enhance the $`\rho \pi `$ branching for $`J/\psi `$ relative to $`\psi (2S)`$. However, the magnitude of $`B(J/\psi \rho \pi )`$ is in line with expectation when we compare the $`B(\rho \pi )/B(\omega \pi )`$ with the inclusive ratio $`B(J/\psi gggX)/B(J/\psi \gamma ^{}X)`$. What happens is not enhancement of $`\rho \pi `$ in $`J/\psi `$ but suppression of $`\rho \pi `$ in $`\psi (2S)`$.<sup>*</sup><sup>*</sup>* More comparison of experiment with models is found in Ref..
Meanwhile the amplitude analysis of the $`J/\psi `$ decay revealed that the relative phase of the gluonic and the one-photon decay amplitude is close to $`90^{}`$ for all two-body decay channels so far studied; $`1^{}0^{}`$, $`0^{}0^{}`$, $`1^{}1^{}`$, and $`N\overline{N}`$. We show in this paper that the recent BES measurement in $`J/\psi 1^+0^{}`$ is also compatible with a large phase.
In contrast, the pattern of a large relative phase does not emerge for $`\psi (2S)`$. Within experimental uncertainties, the relative phase is consistent with zero in the $`1^{}0^{}`$ decay and the $`1^+0^{}`$ decay. This marked difference between $`J/\psi `$ and $`\psi (2S)`$ is another puzzle if the three-gluon decay is equally responsible for the strong decay of $`J/\psi `$ and $`\psi (2S)`$
There is one more experimental information relevant to the issue. That is the hadronic decay rate of $`\psi (2S)`$ which is normally attributed to $`\psi (2S)ggg`$. When we compute with the current data the inclusive gluonic decay rate of $`\psi (2S)`$ by subtracting the cascade and the electromagnetic decay rate from the total rate, it is 60-70% larger, within experimental uncertainties, than what we expect from the short-distance gluonic decay alone. This excess hadronic branching in $`\psi (2S)`$ may suggest that something more occurs in the gluonic decay of $`\psi (2S)`$ than in the $`J/\psi `$ decay.
In this paper we combine these informations together and search the origin of the marked difference between $`J/\psi `$ and $`\psi (2S)`$. While we should be apprehensive about experimental errors at present, they might give us a clue to a solution of the $`\rho \pi `$ puzzle. In Section II, prompted by the experimental observation in the $`0^{}0^{}`$, $`1^{}0^{}`$, $`1^{}1^{}`$, and $`N\overline{N}`$ channels, we postulate universality of the large relative phase between the gluon and the photon decay amplitudes. Specifically, the gluonic decay amplitude acquires a large phase while the photon amplitude is real. We point out that a large phase is consistent with new BES data in $`J/\psi 1^+0^{}`$. Further progress in the BES analysis in this channel will shed more light. We turn to $`\psi (2S)`$ in Section III. The decay branching fractions of $`\psi (2S)1^{}0^{}`$ clearly show suppression of the gluon amplitude and favor a small relative phase between the gluon and the photon amplitude. We point out that a small phase is more likely in $`\psi (2S)1^+0^{}`$ too. Taking the possible excess in the inclusive hadronic decay rate of $`\psi (2S)`$ seriously, we propose that this excess is related to both the suppression and the small relative phase of the $`1^{}0^{}`$ amplitude. Our proposition is that an additional decay process generating the excess should largely cancel the short-distance gluon amplitude in the exclusive decay into $`1^{}0^{}`$ and that the resulting small residual amplitude is not only real but also destructively interferes with the photon amplitude. In Section IV, we first present general consequences of the destructive interference. We then examine two scenarios which may possibly generate the excess inclusive gluonic decay. One is the contribution of the virtual $`D\overline{D}`$ intermediate state. The other is a resonance, a glueball or a four-quark resonance, near the $`\psi (2S)`$ mass. Though neither idea is novel nor highly appealing, they seem to be among a very few possibilities that have not yet been ruled out by experiment.
## II Phases of $`J/\psi `$ decay amplitudes
The relative phase between the gluon and the photon amplitude in the decay $`J/\psi 1^{}0^{}`$ has been analyzed with broken flavor SU(3) symmetry including the $`\rho `$-$`\omega `$ mixing. All analyses clearly show that the relative phase should be very large and not far from $`90^{}`$ with experimetnal uncertainties. The SU(3) analysis was made also for the $`0^{}0^{}`$ modes and the $`1^{}1^{}`$ modes for which leading gluon amplitude is SU(3) violating. The relative phase were found to be equally large for these modes. Furthermore comparison of the electromagnetic form factors in the timelike region with the $`J/\psi `$ decay branching fractions revealed that the relative phase is very close to $`90^{}`$ in the $`N\overline{N}`$ decay channels too. A question arises as to whether this large relative phase is universal to all decays of $`J/\psi `$ or not. There is no persuasive theoretical answer to it at present.Some attempt was recently made to argue in favor of universal large phases.
In addition to those two-body channels already analyzed, the recent BES measurement on $`J/\psi 1^+0^{}`$,
$`B(J/\psi K_1^\pm (1400)K^{})`$ $`=`$ $`(3.8\pm 0.8\pm 1.2)\times 10^3,`$ (1)
$`B(J/\psi K_1^\pm (1270)K^{})`$ $`<`$ $`3.0\times 10^3,\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}90}\%C.L.`$ (2)
is relevant to this issue. We examine here these branching fractions together with the $`b^\pm \pi ^{}`$ branching fraction, $`B(J/\psi b^\pm \pi ^{})=(3.0\pm 0.5)\times 10^3`$.A previous analysis of these $`J/\psi `$ decay modes assumed a zero relative phase and used on a preliminary value of the upper bound on $`B(J/\psi K_1^\pm (1270)K^{})`$. Therefore the $`J/\psi `$ analysis of Ref. should be disregarded. However, the analysis of the $`\psi (2S)1^+0^{}`$ of Ref. remains valid.
Since $`K_1(1270)`$ and $`K_1(1400)`$ are superpositions of $`K_A`$ and $`K_B`$ of the $`1_A^{++}`$ and the $`1_B^+`$ octet, respectively,
$`K_1^+(1400)`$ $`=`$ $`K_A^+\mathrm{cos}\theta +K_B^+\mathrm{sin}\theta ,`$ (3)
$`K_1^+(1270)`$ $`=`$ $`K_A^+\mathrm{sin}\theta +K_B^+\mathrm{cos}\theta ,`$ (4)
with $`\theta 45^{}`$, we can parametrize the three decay amplitudes in terms of the gluon amplitude $`a_1`$ of the $`1_B^+`$ octet and the photon amplitudes $`a_{\gamma A/B}`$ of $`1_A^{++}`$ and $`1_B^+`$;
$`A(b_1^+\pi ^{})`$ $`=`$ $`a_1+\sqrt{1/5}a_{\gamma B},`$ (5)
$`A(K_1^+(1270)K^{})`$ $`=`$ $`(a_1+\sqrt{1/5}a_{\gamma B})\mathrm{cos}\theta a_{\gamma A}\mathrm{sin}\theta ,`$ (6)
$`A(K_1^+(1400)K^{})`$ $`=`$ $`(a_1+\sqrt{1/5}a_{\gamma B})\mathrm{sin}\theta +a_{\gamma A}\mathrm{cos}\theta .`$ (7)
Since there are two independent helicity amplitudes (or $`s`$\- and $`d`$-waves) for $`1^+0^{}`$, we should use this parametrization separately for the $`s`$-wave and the $`d`$-wave amplitudes. The three branching fractions can be fitted with
$`|a_1|`$ $`>`$ $`|a_{\gamma A}||a_{\gamma B}|,`$ (8)
$`\mathrm{arg}(a_1^{}a_{\gamma A})`$ $``$ $`\mathrm{arg}(a_1^{}a_{\gamma B})90^{}.`$ (9)
If the ratio $`\mathrm{\Gamma }(J/\psi \gamma 1^+0^{})/\mathrm{\Gamma }(J/\psi ggg1^+0^{})`$ is comparable in magnitude to $`\mathrm{\Gamma }(J/\psi \gamma X)/\mathrm{\Gamma }(J/\psi gggX)1/5`$, we should expect that $`|a_{\gamma A/B}|0.7|a_1|`$. For $`a_{\gamma A}=a_{\gamma B}=\pm 0.7ia_1`$ and $`\theta 45^{}`$, the ratios of the branching fractions prior to the phase space corrections, denoted by $`B_0`$, take values as
$$B_0(b^\pm \pi ^{}):B_0(K_1^\pm (1270)K^{}):B_0(K_1^\pm (1400)K^{})1:0.5:0.9.$$
(10)
While the inequality $`B(K_1^\pm (1270)K^{})<B(K_1^\pm (1400)K^{})`$ can easily be realized by a wide range of parameter values, the other inequality $`B(b^\pm \pi ^{})<B(K_1^\pm (1400)K^{})`$ is a little tight. If we allow $`|a_{\gamma A/B}|`$ larger than $`0.7|a_1|`$ and/or increase the value of $`\theta `$, however, the current central values of the branching fractions are consistent with the large phase hypothesis. We should also point out that if the SU(3) breaking correction is made by the meson wavefunctions $`(f_Pf_A)^2`$, it is likely to enhance $`B(K_1^\pm (1400)K^{})`$ over $`B(b^\pm \pi ^{})`$ to the direction in favor of the large phase fit.
If we leave $`a_{\gamma A/B}`$ unrestricted in magnitude and phase, a triangular relation holds for the amplitudes as
$$A(K_1^\pm (1270)K^{})\mathrm{cos}\theta +A(K_1^\pm (1400)K^{})\mathrm{sin}\theta =A(b^\pm \pi ^{}).$$
(11)
Determination of the $`s`$-to-$`d`$ wave ratio of the amplitudes and further study of $`B(J/\psi K_1^\pm (1400)K^{})`$ will eventually resolve the composition of amplitudes and test the large phase hypothesis in $`1^+0^{}`$. As it was pointed out previously, it is also important to resolve the discrepancy between $`B(b^\pm \pi ^{})`$ and $`2B(b^0\pi ^0)`$, which theory predicts to be equal.
To summarize the experimental situation of the two-body $`J/\psi `$ decay amplitudes, the existing data strongly favor large relative phases close to $`90^{}`$ between the gluon and the photon decay amplitudes for $`1^{}0^{}`$, $`0^{}0^{}`$, $`1^{}1^{}`$, and $`N\overline{N}`$, and are consistent with a large phase for $`1^+0^{}`$.
What does theory say about these relative phases ? In the perturbative picture, the gluonic decay of $`J/\psi `$ proceeds as depicted in Fig.1a. The inclusive decay rate is computed with the gluons placed on mass shell. In contrast, the photon being far off shell, no corresponding on-shell intermediate state appears in the perturbative diagrams of the photon amplitude (Fig.1b). Although perturbative QCD (pQCD) is a good description of inclusive charmonium decays, it is questionable whether it works for two-body decay channels of charmonia. To be specific, the pQCD prediction of the asymptotic pion form factor,
$$F_\pi (q^2)16\pi \alpha _s(q^2)f_\pi ^2/q^2,$$
(12)
has not been reached at the $`J/\psi `$ mass. Furthermore, the helicity suppression argument of pQCD fails for the $`\omega \pi `$ decay channel. Though it is tempting, therefore, we cannot argue for the large relative phases on the basis of perturbative diagrams. Whether or not these large relative phases are universal to all two-body decay modes of $`J/\psi `$ must be determined by experiment. Despite lack of a good theoretical argument at present, we suspect nonetheless that the universal large phases so far found are not an accident.
If the relative phases are close to $`90^{}`$, it is more likely that the photon decay amplitudes are real and consequently the gluon decay amplitudes are imaginary. The reason is as follows: In order for the photon decay amplitude to have a substantial phase, the final $`q\overline{q}`$ created by the virtual $`\gamma `$ should have a large absorptive part of a long-distance origin. This can happen if there should be a relatively sharp $`q\overline{q}`$ resonance just around $`J/\psi `$. More likely is that many resonances exist below $`J/\psi `$ as in the vector-meson-dominance scenario, or that as in the dual resonance model, many increasingly broader resonances appear all the way to high energies. In the former case, only the tails of low-lying resonances contribute to the real part. In the latter case, a few nearby broad resonances can contribute to the imaginary part, but they are outnumbered by many more resonances below and above $`J\psi `$ that contribute to the real part. In comparison, we have less insight in hadronization dynamics of the gluon decay.
Motivated by the results in the amplitude analyses of the two-body of $`J/\psi `$ decays, we make two postulates:
(1) The relative phases between the gluon and the photon decay amplitudes are universally large for all two-body decays of $`J/\psi `$. The photon decay amplitudes are predominantly real and consequently the gluon decay amplitudes are imaginary.
(2) The same pattern holds for $`\psi (2S)`$ decay as well.
These are the starting assumptions of our analysis that follows.
## III $`\psi (2S)`$ decay
### A Relative phase from experiment
The only large energy scale involved in the three-gluon decay of charmonia is the charm quark mass $`m_c`$. Whether one accepts the argument of the universal large phase in exclusive channels or not, therefore, one would naively expect that the corresponding phases should not be much different between the $`J/\psi `$ decay and the $`\psi (2S)`$ decay. However, experimental data so far available show that the phases are small at least in some two-body decay modes of $`\psi (2S)`$. The strongest evidence is in the decay $`\psi (2S)1^{}0^{}`$, which includes the puzzling $`\psi (2S)\rho \pi `$. In the case that the final $`0^{}`$ meson is an octet, we can parametrize the $`\psi (2S)1^{}0^{}`$ decay amplitudes with the SU(3) singlet amplitude $`a_1`$, the SU(3) breaking correction $`ฯต`$ due to $`m_sm_{ud}`$, and the photon amplitude $`a_\gamma `$. The corresponding amplitudes, $`b_1`$ and $`b_\gamma `$, are introduced for the $`0^{}`$ singlet. For the $`\varphi `$-$`\omega `$ mixing, we assume the nonet scheme. The parametrization of the amplitudes is listed in Table I for the decay modes so far studied in experiment. Since the analyzed channels are limited and the uncertainties in their branching fractions are still large, we are unable to perform a meaningful $`\chi ^2`$ fit at present. Therefore, we present only fits to the central values by referring to Table I.
First of all, if we ignored the photon amplitude $`a_\gamma `$, we would obtain $`B(K^0\overline{K}^0+c.c.)=B(K^\pm K^{})`$, which contradicts with experiment, $`(0.81\pm 0.24\pm 0.16)\times 10^4`$ vs $`<0.30\times 10^4`$. The large splitting between these branching fractions requires that $`a_\gamma `$ be comparable to $`a_1`$. If we set $`a_1`$ and $`ฯต`$ to zero, we would obtain up to phase space corrections
$$B_0(\omega \pi )/B_0(K^0\overline{K}^0+c.c.)=9/8$$
(13)
in contradiction with the measurement, $`(0.38\pm 0.17\pm 0.11)/(0.81\pm 0.24\pm 0.16)`$.<sup>ยง</sup><sup>ยง</sup>ยง It is possible that the SU(3) breaking in the strange and nonstrange meson wavefunctions $`(f_\pi f_\rho /f_Kf_K^{})^2`$ may be responsible for part of the discrepancy. In order to come closer to this ratio of the measured values, a large constructive interference should occur in $`K^0\overline{K}^0`$, that is, the relative phase must be small between $`a_1+ฯต`$ and $`2a_\gamma `$. Then, assuming that $`a_1`$ and $`ฯต`$ have a common phase, we have a large destructive interference between $`a_1`$ and $`a_\gamma `$ for both $`\rho \pi `$ and $`K^\pm K^{}`$ in agreement with experiment. This main feature of the fit to the $`\psi (2S)1^{}0^{}`$ amplitudes is found in the earlier paper by Chen and Braaten and, in particular, in the paper by Tuan. This solves the $`\rho \pi `$ puzzle and explains also the missing of the $`K^\pm K^{}`$ mode in experiment.
With these qualitative observations in mind, we have fitted to the central values of the observed branching fractions and then have computed with those parameter values the branching fractions of the modes for which only the upper bounds have been determined. In Table I we have listed the fit with $`\delta \mathrm{arg}(a_1^{}a_\gamma )=0^{}`$ and the large phase fit with $`\delta =\pm 90^{}`$ for comparison. When the central values of $`B(\omega \pi )`$ and $`B(K^0\overline{K}^0+c.c.)`$ are fitted with $`\delta =0^{}`$, the ratio of the photon and the gluon amplitude turns out to be
$$a_\gamma /(a_1+ฯต)0.76.$$
(14)
For comparison, $`a_\gamma /(a_1+ฯต)0.14`$ in the case of $`J/\psi `$. We would expect $`|a_\gamma /a_1|0.22`$ if the ratio of $`\mathrm{\Gamma }(\psi (2S)\gamma ^{}1^{}0^{})`$ to $`\mathrm{\Gamma }(\psi (2S)ggg1^{}0^{})`$ is roughly equal to $`\mathrm{\Gamma }(\psi (2S)\gamma ^{}X)/\mathrm{\Gamma }(\psi (2S)gggX)`$. Since experiment shows that $`\mathrm{\Gamma }(\omega \pi )/\mathrm{\Gamma }(l^+l^{})(|a_\gamma |^2)`$ is about the same for $`J/\psi `$ and $`\psi (2S)`$, the large number in Eq.(14) results from strong suppression of the total gluonic amplitude $`a_1+ฯต`$ in $`\psi (2S)`$. As the value of $`ฯต`$ is varied in the range of $`|ฯต/a_1|<1/3`$, the value of $`B(\rho ^\pm \pi ^{})`$ varies between 0 and 0.04$`\times 10^4`$. The values for $`B(\rho ^\pm \pi ^{})`$ and $`B(K^\pm K^{})`$ can be increased if we stretch within the experimental uncertainties of $`B(\omega \pi )`$ and $`B(K^0\overline{K}^0+c.c.)`$. In contrast, the fit with $`\delta =\pm 90^{}`$ overshoots the upper bound on $`B(K^\pm K^{})`$ and, if $`|ฯต|<\frac{1}{3}|a_1|`$, the upper bound on $`B(\rho ^\pm \pi ^{})`$ very badly. A fit with $`\delta =\pm 90^{}`$ is virtually impossible even with experimental uncertainties unless $`|ฯต||a_1|`$. We thus conclude that the relative phase between $`a_1`$ and $`a_\gamma `$ should be small in $`\psi (2S)1^{}0^{}`$ contrary to the $`J/\psi `$ decay.
Though it is less conclusive, a small phase seems to be favored in the $`1^+0^{}`$ decay of $`\psi (2S)`$ too. It is conspicuous in experiment that the $`K_1^\pm (1400)K^{}`$ mode is strongly suppressed relative to the $`K_1^\pm (1270)K^{}`$:
$`B(\psi (2S)K_1^\pm (1270)K^{})`$ $`=`$ $`(10.0\pm 1.8\pm 2.1)\times 10^4,`$ (15)
$`B(\psi (2S)K_1^\pm (1400)K^{})`$ $`<`$ $`3.1\times 10^4.`$ (16)
$`B(\psi (2S)b^\pm \pi ^{})`$ is half way between them:
$$B(\psi (2S)b^\pm \pi ^{})=(5.2\pm 0.8\pm 1.0)\times 10^4.$$
(17)
We can use Eq. (4) as the parametrization of $`\psi (2S)1^+0^{}`$. First of all, if $`a_1`$ dominated over $`a_{\gamma A/B}`$, we would have $`B(b^\pm \pi ^{})2B(K_1^\pm (1270)K^{})2B(K_1^\pm (1400)K^{})`$ for $`\theta 45^{}`$ in disagreement with experiment. Just as in $`\psi (2S)1^{}0^{}`$, $`|a_1|`$ is comparable to $`|a_{\gamma A/B}|`$. Next, the strong suppression of $`K_1^\pm (1400)K^{}`$ relative to $`K_1^\pm (1270)K^{}`$ can be realized only when $`a_1+\sqrt{1/5}a_{\gamma B}`$ interferes destructively with $`a_{\gamma A}`$. Therefore, the relative phase between $`a_1`$ and $`a_{\gamma A/B}`$ must be small modulo $`\pi `$. To obtain $`B(K_1^\pm (1270)K^{})2B(b^\pm \pi ^{})`$, we need $`a_1+\sqrt{1/5}a_{\gamma B}a_{\gamma A}`$. The allowed range of the amplitude ratios was plotted in Ref. by choosing all amplitudes as relatively real and assuming tentatively the s-wave decay for phase-space corrections. Though it is not impossible to fit the three branching fractions with $`\theta 90^{}`$, we must have $`a_10`$ and $`\sqrt{1/5}a_{\gamma B}a_{\gamma A}`$ in that case.
To summarize for $`\psi (2S)`$, the data on $`1^{}0^{}`$ virtually excludes the possibility of a small phase between $`a_1`$ and $`a_\gamma `$. A fit to $`1^+0^{}`$ has more room when the relative phase is small. There is no evidence for that the relative phase must be large in $`\psi (2S)`$. For some reason, a large relative phase does not seem to occur in the $`\psi (2S)`$ decay. We ask what causes this marked difference between $`J/\psi `$ and $`\psi (2S)`$ when we postulate the universal large phase for $`\psi (2S)`$ as well as for $`J/\psi `$.
### B Excess hadronic rate in inclusive hadronic decay
It has been noticed that when one computes the inclusive hadronic decay rate of $`\psi (2S)`$ through $`ggg`$ by subtracting the rates of the cascade and electromagnetic decays from the total decay rate, it is substantially larger than what we expect from an extrapolation of $`J/\psi `$. The number with a conservative error estimate based on the listings of Reviews of Particle Physics is
$$\frac{B(\psi (2S)ggg+gg\gamma )}{B(J/\psi ggg+gg\gamma )}=0.23\pm 0.07,$$
(18)
which should be compared with
$$\left(\frac{\alpha _s(\psi (2S))}{\alpha _s(J/\psi )}\right)^3\frac{B(\psi (2S)ล^+ล^{})}{B(J/\psi ล^+ล^{})}=0.134\pm 0.034.$$
(19)
Smaller errors ($`0.226\pm 0.052`$ vs $`0.141\pm 0.012`$) have been attached in a recent literature with a different error estimate. We would expect that the two numbers should be equal to each other since the wavefunctions at origin appear in common in Eqs.(18) and (19). The discrepancy of 60-70% between them alarms us particularly because all numbers involved have been repeatedly measured over many years.The author learned that BES collaboration is considering a different determination of the cascade decay branchings.
In comparison we find no similar excess in $`\mathrm{{\rm Y}}(2S)`$ though experimetnal undertaintues are large. In terms of the ratio of branching ratios, $`\overline{B}(ggg+gg\gamma )B(\mathrm{{\rm Y}}ggg+gg\gamma )/\alpha _s(\mathrm{{\rm Y}})^3B(\mathrm{{\rm Y}}\mu ^+\mu ^{})`$, three $`\mathrm{{\rm Y}}`$โs are more in line:
$$\overline{B}(ggg+gg\gamma )=\{\begin{array}{cc}(4.5\pm 0.2)\times 10^3,& \mathrm{{\rm Y}}(1S)\\ (4.9\pm 0.9)\times 10^3,& \mathrm{{\rm Y}}(2S)\\ (4.0\pm 0.4)\times 10^3,& \mathrm{{\rm Y}}(3S).\end{array},$$
(20)
where the total leptonic branching for $`\mathrm{{\rm Y}}(3S)`$<sup>\**</sup><sup>\**</sup>\** The branching to $`\mathrm{{\rm Y}}(2S)\gamma \gamma `$ quoted in is sum of the cascade decay branchings $`\chi _{bJ}(2P)\gamma \mathrm{{\rm Y}}(2S)\gamma `$ in view of the $`\gamma \gamma `$ invariant mass spectrum and also of its magnitude ($`>B(\mathrm{{\rm Y}}(2S)\pi ^0\pi ^0)`$). It is counted in the radiative decay branchings separately listed in . has been substituted with three times $`B(\mu ^+\mu ^{})`$, the only quoted leptonic branching. It appears that the excess in $`\overline{B}(ggg+gg\gamma )`$ is unique to $`\psi (2S)`$. However, this excess in the inclusive rate has not shown up in the rates of the exclusive channels so far measured. In fact, the ratio $`B(\psi (2S)h)/B(J/\psi h)`$ scatter around the expected value ($``$ 13-14% of Eq.(19)), which was often called the 14% rule. Some remarks should be in order on it.
First of all, the 14% rule is largely violated in many of two-body and quasi-two-body channels, as we recently learned in the BES data. The $`\rho \pi `$ channel is an extreme case. For multihadron channels, there are actually not so many modes that are available for testing the 14% rule.<sup>โ โ </sup><sup>โ โ </sup>โ โ Gu and Li lumped multihadron modes together and compared between $`J/\psi `$ and $`\psi (2S)`$. Then the number is dominated by three modes, $`\pi ^+\pi ^{}\pi ^0`$, $`2(\pi ^+\pi ^{})\pi ^0`$ and $`3(\pi ^+\pi ^{})\pi ^0`$, which happen to be below 14%. The mode $`\pi ^+\pi ^{}\pi ^0`$ is actually $`\rho \pi `$ and its nonresonant content is consistent with zero. More revealing are the ratios of individual modes. In Table II, we have tabulated the ratios for the modes not listed in but available for comparison. We see that the ratios scatter rather widely above and below 14% with some tendency of being smaller than 14%, but with fairly large experimental uncertainties. It is important to notice that the branching fractions of all modes in Table II add up to no more than 15% of the total gluonic decay branching of $`J/\psi `$. Indeed, only one charge state has been available for comparison from each of $`5\pi `$, $`7\pi `$, $`2\pi K\overline{K}`$, and $`N\overline{N}n\pi `$. We have not yet seen comparison of the rest. The so-called 14% rule is based on very limited number of decay modes. It is premature to preclude the hadronic excess with the data of multihadron exclusive channels.
If future experiment shows that the hadronic excess in $`\psi (2S)`$ is real, it may have something to do with the $`\rho \pi `$ puzzle and with the abrupt change of the relative phase of amplitudes from $`J/\psi `$ to $`\psi (2S)`$. A process responsible for the excess inclusive hadron rate can interfere with the short-distance gluon decay in exclusive modes. If its amplitude makes a large destructive interference with the three-gluon amplitude in $`\psi (2S)1^{}0^{}`$ and if the sum is nearly real and comparable to the photon amplitude in magnitude, our puzzle can be solved. We shall look into possible sources of this rate excess in the following.
## IV Additional hadronic amplitude in $`\psi (2S)`$
### A General consequences
If the origin of the problem is in the interference of an unknown additional process with the short-distance gluon decay of $`\psi (2S)`$, we expect a general pattern of correlation between the decay angular distribution and suppression or enhancement.
The decay angular distribution for two final hadrons is generally of the form,
$$d\mathrm{\Gamma }/d\mathrm{\Omega }1+a\mathrm{cos}^2\theta ,(|a|1)$$
(21)
where $`\theta `$ is the polar angle measured from the $`e^+e^{}`$ beam direction. For $`1^{}0^{}`$ and $`0^{}0^{}`$ decays, the value of $`a`$ is constrained kinematically to +1 and $`1`$, respectively, while it is determined dynamically by the helicity content, $`\pm 1`$ or 0, of the final state in other decays. In $`1^{}0^{}`$ and $`0^{}0^{}`$, therefore, any additional amplitude has the same angular dependence as the three-gluon and the photon amplitude irrespective of its origin. Consequently a high degree of destructive or constructive interference with an additional amplitude is possible in these decays. Observation of the strongest suppression in the $`1^{}0^{}`$ mode is consistent with this pattern. We expect that the decay rates of $`\psi (2S)0^{}0^{}`$ may also be quite different from those of $`J/\psi 0^{}0^{}`$. In terms of $`\overline{B}(0^{}0^{})B(0^{}0^{})/\alpha _s^3B(\mu ^+\mu ^{})`$, the current data give
$`\overline{B}(\pi ^+\pi ^{})`$ $`=`$ $`\{\begin{array}{cc}0.15\pm 0.02\hfill & \text{for }J/\psi \hfill \\ 0.8\pm 0.5\hfill & \text{for }\psi (2S)\hfill \end{array}`$ (24)
$`\overline{B}(K^+K^{})`$ $`=`$ $`\{\begin{array}{cc}0.24\pm 0.03& \text{for }J/\psi \\ 0.94\pm 0.66& \text{for }\psi (2S)\end{array}`$ (27)
Within the large experimental uncertainties we see a hint of large constructive interference in $`\psi (2S)0^{}0^{}`$. In contrast, in other processes an additional amplitude and the three-gluon amplitude have different angular distributions in general. A large interference can occur only when dynamical mechanisms of two processes are similar. Otherwise it should be a result of a high degree of accident. When a large disparity is observed between corresponding two-meson decay rates of $`J/\psi `$ and $`\psi (2S)`$, therefore, the decay angular distribution of this channel will also be very different between $`J/\psi `$ and $`\psi (2S)`$. This will give a good test of the idea of interference with an additional amplitude.
The other consequence is in multibody final states. Since the additional process enhances the inclusive rate, a large number of exclusive decay channels should receive enhancement rather than suppression. When there are many hadrons in the final state, chance of interference between amplitudes of different decay mechanism is much smaller because of difference in subenergy dependence and event topology. Therefore enhancement will not be dramatic. While we have not yet seen such enhancemnet in Table II, we expect that the branching fraction tend to be enhanced in many nonresonant multibody channels of $`\psi (2S)`$ relative to $`J/\psi `$.
Where does the additional amplitude possibly come from ? There are few options left in modifying charmonium physics radically. Since aspects of perturbative QCD have been well understood, we are bound to look for the origin of the problem in long-distance physics of one kind or another.
### B $`\psi (2S)D\overline{D}`$ hadrons
One unique feature of $`\psi (2S)`$ is a close proximity of its mass to the $`D\overline{D}`$ threshold. The $`\psi (2S)`$ mass is only 43 MeV (53 MeV) below $`D^0\overline{D}^0`$ ($`D^+D^{}`$), while $`\mathrm{{\rm Y}}(3S)`$ is 200 MeV away from the $`B\overline{B}`$ threshold. Can the small energy difference<sup>โกโก</sup><sup>โกโก</sup>โกโกThis was brought to the authorโs attention by J.L. Rosner. between $`\psi (2S)`$ and $`D\overline{D}`$ have anything to do with the excess ? It may happen that $`\psi (2S)`$ picks up a light quark pair through soft gluons and dissociates virtually into $`D\overline{D}`$, which in turn annihilate into light hadrons. (See Fig.2.) The dominant process of the $`D\overline{D}`$ annihilation is through $`c\overline{c}`$ annihilation through a single hard gluon. The small energy denominator enhances creation of virtual $`D\overline{D}`$ while $`p`$-wave creation compensates the enhancement. It is difficult, actually nearly impossible, to give a reliable computation of this sequence. Deferring estimate of the rate to future, we shall comment here only on whether the $`D\overline{D}`$ contribution can have a final-state interaction phase large enough to cancel the perturbative gluon amplitude or not.
Since the $`D\overline{D}`$ intermediate state is above the $`\psi (2S)`$ mass, a phase of amplitude must come from the subsequent annihilation of $`c\overline{c}`$ and thereafter. After an energetic light quark pair $`q\overline{q}`$ is created from $`c\overline{c}`$, each of $`q\overline{q}`$ picks up a soft light quark from the light quark cloud of $`D\overline{D}`$ to form mesons. (See Fig.3a.) Kinematically, this step of the hadron formation process is quite different from that of the timelike electromagnetic form factor of a meson in which energetic light quarks pick up collinear quarks created by a hard gluon. (See Fig.3b.) In our case color-dipole moment is large for all pairs of quarks. Furthermore, the c.m. energy of a hard quark in one meson and a soft quark in the other meson is in the low energy resonance region,
$$\sqrt{s}=O(\sqrt{2\mathrm{\Lambda }_{QCD}m_c})<1\mathrm{G}\mathrm{e}\mathrm{V}.$$
(28)
Therefore, one cannot argue that final-state interactions should be small between final mesons. We expect that there is a good chance for the amplitude of $`\psi (2S)D\overline{D}mesons`$ to acquire a substantial final-state interaction phase. Weakness in this argument is that the phase can be large but need not be large.
The idea of the virtual $`D\overline{D}`$ dissociation actually has some common feature with that of the โhigher Fock componentโ of charmonia. The $`D\overline{D}`$ state can be viewed as part of the four-quark Fock space of $`\psi (2S)`$. The higher Fock component was proposed as an additional contribution to $`J/\psi 1^{}0^{}`$ to solve the $`\rho \pi `$ puzzle. It was argued that it is more significant in $`J/\psi `$ than in $`\psi (2S)`$. As we have emphasized, however, there is nothing anomalous about $`J/\psi 1^{}0^{}`$.
### C $`\psi (2S)`$resonance$``$ hadrons
The second idea is a twist of an old one: A noncharm resonance may exist near the $`\psi (2S)`$ mass and give an extra contribution to the hadronic decay rate. A glueball was proposed earlier at the $`J/\psi `$ mass to boost the $`\rho \pi `$ decay rate of $`J/\psi `$. However, we now want it near $`\psi (2S)`$ not near $`J/\psi `$. We look into the possibility that some resonance around the $`\psi (2S)`$ mass, a glueball or four-quark, destructively interferes with the perturbative $`\psi (2S)ggg1^{}0^{}`$ decay. Admittedly, the idea is ad hoc and there is some difficulty aside from unnaturalness.
Light-quark resonances of high mass ($`3.7`$ GeV) and low spin are normally too broad to be even recognized as resonances. Four-quark resonances may be an alternative if they exist at all. The mass of $`3.7`$ GeV is normally considered as too high for the lowest vector glueball. An excited glueball state of $`J^PC=1^{}`$ serves our purpose. Whatever its origin is, let us introduce here such a resonance, call it $`R`$, and see its consequences.
In order for $`\psi (2S)`$ to decay through $`R`$ as strongly as through three gluons, the coupling $`f`$ of $`R`$ to $`\psi (2S)`$ defined by $`fm_R\psi _\mu R^\mu `$ must be large enough. The $`\psi (2S)`$-$`R`$ mixing at the $`\psi (2S)`$ mass is given by
$$\epsilon \frac{f}{\mathrm{\Delta }mi\mathrm{\Gamma }_R},$$
(29)
where $`\mathrm{\Delta }m=m_Rm(\psi (2S))`$ and $`\mathrm{\Gamma }_R`$ is the total width of $`R`$. It leads to $`\mathrm{\Gamma }(\psi (2S)R\mathrm{hadrons})|f|^2/\mathrm{\Gamma }_R`$ when $`|\mathrm{\Delta }m|<O(\mathrm{\Gamma }_R)`$. To obtain $`\mathrm{\Gamma }(\psi (2S)R\mathrm{hadrons})\mathrm{\Gamma }(\psi (2S)ggg)`$, we need therefore
$$|f|^2\mathrm{\Gamma }_R\mathrm{\Gamma }(\psi (2S)ggg).$$
(30)
If $`R`$ is a light-quark resonance $`q\overline{q}`$, $`|f|`$ would be much too small for the following reason: While $`|f|^2`$ is of the order of $`\mathrm{\Gamma }(\psi (2S)ggg)\mathrm{\Gamma }(Rggg)`$ for $`q\overline{q}`$, we expect $`\mathrm{\Gamma }(Rggg)\mathrm{\Gamma }_R`$ because of the $`\alpha _s^3`$ suppression of $`q\overline{q}ggg`$. Therefore there is no chance to satisfy Eq.(30). It is likely that the same argument applies to four-quark resonances.
For glueballs, we simply do not have enough quantitative understanding to rule out a large enough coupling to $`\psi (2S)`$. Hou and Soni proposed a glueball near the $`J/\psi `$ mass in order to enhance $`J/\psi \rho \pi `$ (and its symmetry-related modes) but not other decay modes. To accomplish it, this glueball must have very special, if not unnatural, properties: It is nearly degenerate with $`J/\psi `$ with a quite narrow width for an object of mass $`3`$ GeV and decays predominantly into $`1^{}0^{}`$. Later Hou relaxed the constraint on the $`\rho \pi `$ branching to argue that such a glueball was not yet ruled out by the search of the BES Collaboration.
In our case, since a glueball is introduced to account for the hadronic excess, it should couple not primarily to the $`1^{}0^{}`$ channels, but to many other channels. What we need is a generic vector glueball with no special or unusual properties. If $`\mathrm{\Gamma }_R`$ is as narrow as 100 MeV, for instance, the mixing $`|\epsilon |=O(10^2)`$ would be able to account for the excess in the inclusive hadron decay of $`\psi (2S)`$. The width can be wider. In that case the pole transition strength $`f`$ should be stronger according to Eq.(30). Since the glueball $`R`$ couples to a photon only indirectly through its mixing to a quark pair, it is hard to detect $`R`$ in the hadronic cross section of $`e^+e^{}`$ annihilation near the $`\psi (2S)`$ mass. Searching by hadronic reactions such as $`p\overline{p}`$ annihilation is a daunting task. From a purely experimental viewpoint, such a resonance has not been ruled out.
Though the resonance scenario is admittedly a long shot, it is one of a very few options left to us. One reason to pursue this somewhat unnatural scenario is that the amplitude for $`\psi (2S)RX`$ has automatically a large phase when $`\mathrm{\Delta }m<O(\mathrm{\Gamma }_R)`$ since the coupling $`f`$ is likely real, that is, dominated by the dispersive part. If that is the case, the resonant amplitude can interfere strongly with the three-gluon amplitude in two-meson decays.
## V Conclusion
We have searched for a clue to solve the $`\rho \pi `$ puzzle in this paper. Our purpose is to locate the source of the problem rather than to offer a final solution of the problem. Two threads have been exposed which may eventually lead us to a solution of the $`\rho \pi `$ puzzle. They are the phases of the decay amplitudes and a possible excess in the inclusive hadronic decay rate of $`\psi (2S)`$. An experimental confirmation of the excess will be the most useful in directing theorists. If it is confirmed, it will be quite an important experimental discovery by itself. One crucial experimental information will be the angular distributions of $`J/\psi `$ and $`\psi (2S)`$ into the channels other than $`1^{}0^{}`$ and $`0^{}0^{}`$. Difference in the angular distributions should have direct correlation with enhancement and suppression in general. As for the source of an additional process, the virtual $`D\overline{D}`$ pair and the vector glueball are two options that cannot be ruled out. To be frank, however, we admit that both ideas have unnaturalness. More an attractive alternative is highly desired. It is possible that the large relative phases so far observed in $`J/\psi `$ decay are an accident and that the $`\rho \pi `$ puzzle is a problem of incalculable long-distance complications. However, our hope is that there might be something novel, simple, or fundamental hidden beneath the issue.
###### Acknowledgements.
The author thanks F.A. Harris and D. Paluselli for communications about the BES analysis and J.L. Rosner for a stimulating suggestion. This work was supported in part by the Director, Office of Science, Office of High Energy and Nuclear Physics, of the U.S. Department of Energy under Contract DEโAC03โ76SF00098 and in part by the U.S. National Science Foundation under grant PHYโ95โ14797.
|
warning/0006/hep-th0006051.html
|
ar5iv
|
text
|
# Dynamical Symmetry Breaking and Magnetic Confinement in QCD
## Abstract
We present a gauge independent method to construct the effective action of QCD, and calculate the one loop effective action of $`SU(2)`$ QCD in an arbitrary constant background field. Our result establishes the existence of a dynamical symmetry breaking by demonstrating that the effective potential develops a unique and stable vacuum made of the monopole condensation in one loop approximation. This provides a strong evidence for the magnetic confinement of color through the dual Meissner effect in the non-Abelian gauge theory. The result is obtained by separating the topological degrees which describe the non-Abelian monopoles from the dynamical degrees of the gauge potential, and integrating out all the dynamical degrees of QCD. We present three independent arguments to support our result.
## I Introduction
One of the most outstanding problems in theoretical physics is the confinement problem in QCD. It has long been argued that the monopole condensation could explain the confinement of color through the dual Meissner effect . Indeed, if one assumes the monopole condensation, one could easily argue that the ensuing dual Meissner effect guarantees the confinement . In this direction there has been a remarkable progress in the lattice simulation during the last decade. In fact the recent numerical simulations have provided an unmistakable evidence which supports the idea of the magnetic confinement through the monopole condensation . Unfortunately so far there has been no satisfactory field theoretic proof of the monopole condensation in QCD. The purpose of this paper is to present a gauge independent method to construct the effective action of QCD, and to establish the magnetic confinement in the non-Abelian gauge theory from the first principles of the quantum field theory . Utilizing a gauge independent parameterization of the gluon potential which emphasizes its topological character we establish the existence of the non-trivial vacuum made of the monopole condensation in SU(2) QCD in one loop approximation, after integrating out all the dynamical degrees of the non-Abelian potential. Our analysis shows that it is precisely the magnetic moment interaction of the gluons which was responsible for the asymptotic freedom that generates the monopole condensation in QCD. This strongly indicates that the magnetic confinement is indeed the correct confinement mechanism of color in QCD.
To prove the magnetic confinement it is instructive for us to remember how the magnetic flux is confined in the superconductor through the Meissner effect. In the macroscopic Ginzburg-Landau description of superconductivity the Meissner effect is triggered by the effective mass of the electromagnetic potential, which determines the penetration (confinement) scale of the magnetic flux. In the microscopic BCS description, this effective mass is generated by the electron-pair (the Cooper pair) condensation. This suggests that, for the confinement of the color electric flux, one needs the condensation of the monopoles. Equivalently, in the dual Ginzburg-Landau description, one needs the dynamical generation of the effective mass for the monopole potential. To demonstrate this one must first identify the monopole potential, and separate it from the generic QCD connection, in a gauge independent manner. This can be done with an โAbelianโ projection , which provides us a natural reparameterization of the non-Abelian connection in terms of the restricted connection (i.e., the dual potential) of the maximal Abelian subgroup $`H`$ of the gauge group $`G`$ and the valence gluon (i.e., the gauge covariant vector field) of the remaining $`G/H`$ degrees. With this separation one can show that the monopole condensation takes place in one loop correction, after one integrates out all the dynamical degrees of the non-Abelian gauge potential.
The monopole condensation by itself does not guarantee that it describes the true vacuum of QCD. To prove that the monopole condensate does indeed describe the true vacuum, one must calculate the effective potential with an arbitrary background field configuration and show that the monopole condensate becomes the absolute minimum of the effective potential. In the following we prove that this is indeed the case, at least in one loop approximation. We show that with an arbitrary background the color electric field creates an instability to the effective action by generating an imaginary part. This proves that the monopole condensation provides the only stable vacuum of QCD which is unique. As importantly our analysis shows that the gluon loop contributes positively, but the quark loop contributes negatively, to the imaginary part of the effective action. This means that the gluons generate an anti-screening effect by making pair annihilations, while the quarks generate a screening effect by making pair creations. This is a very important observation, because this indicates that the gluons are not able to form a hadronic bound state. A big mystery in hadron spectroscopy has been the absence of the glueball states made of the valence gluons. Our analysis provides a natural explanation why this is so.
## II Abelian Projection and Extended QCD
Consider $`SU(2)`$ QCD for simplicity. A natural way to identify the monopole potential is to introduce an isotriplet unit vector field $`\widehat{n}`$ which selects the โAbelianโ direction (i.e., the color charge direction) at each space-time point, and to decompose the connection into the restricted potential $`\widehat{A}_\mu `$ which leaves $`\widehat{n}`$ invariant and the valence gluon $`\stackrel{}{X}_\mu `$ which forms a covariant vector field ,
$`\stackrel{}{A}_\mu =A_\mu \widehat{n}{\displaystyle \frac{1}{g}}\widehat{n}\times _\mu \widehat{n}+\stackrel{}{X}_\mu =\widehat{A}_\mu +\stackrel{}{X}_\mu ,`$ (1)
$`(\widehat{n}^2=1,\widehat{n}\stackrel{}{X}_\mu =0),`$ (2)
where $`A_\mu =\widehat{n}\stackrel{}{A}_\mu `$ is the โelectricโ potential. Notice that the restricted potential is precisely the connection which leaves $`\widehat{n}`$ invariant under the parallel transport,
$`\widehat{D}_\mu \widehat{n}=_\mu \widehat{n}+g\widehat{A}_\mu \times \widehat{n}=0.`$ (3)
Under the infinitesimal gauge transformation
$`\delta \widehat{n}=\stackrel{}{\alpha }\times \widehat{n},\delta \stackrel{}{A}_\mu ={\displaystyle \frac{1}{g}}D_\mu \stackrel{}{\alpha },`$ (4)
one has
$`\delta A_\mu ={\displaystyle \frac{1}{g}}\widehat{n}_\mu \stackrel{}{\alpha },\delta \widehat{A}_\mu ={\displaystyle \frac{1}{g}}\widehat{D}_\mu \stackrel{}{\alpha },`$ (5)
$`\delta \stackrel{}{X}_\mu =\stackrel{}{\alpha }\times \stackrel{}{X}_\mu .`$ (6)
This shows three things. First the restricted potential by itself forms an $`SU(2)`$ connection which satisfies the full $`SU(2)`$ gauge degrees of freedom. Moreover, $`\widehat{A}_\mu `$ retains the full topological characteristics of the original non-Abelian potential. Clearly the isolated singularities of $`\widehat{n}`$ define $`\pi _2(S^2)`$ which describes the non-Abelian monopoles. Indeed $`\widehat{A}_\mu `$ with $`A_\mu =0`$ and $`\widehat{n}=\widehat{r}`$ describes precisely the Wu-Yang monopole . Besides, with the $`S^3`$ compactification of $`R^3`$, $`\widehat{n}`$ characterizes the Hopf invariant $`\pi _3(S^2)\pi _3(S^3)`$ which describes the topologically distinct vacua . Secondly the valence gluon forms a gauge covariant colored source of the restricted potential which does not inherit any non-linear characters of the non-Abelian connection. Finally this decomposition of the non-Abelian connection is made without compromising the gauge invariance. Obviously the decomposition holds in any gauge, and is gauge independent.
The above discussion tells that $`\widehat{A}_\mu `$ has a dual structure. Indeed the field strength made of the restricted potential is decomposed as
$`\widehat{F}_{\mu \nu }=(F_{\mu \nu }+H_{\mu \nu })\widehat{n}\text{,}`$ (7)
$`F_{\mu \nu }=_\mu A_\nu _\nu A_\mu ,H_{\mu \nu }={\displaystyle \frac{1}{g}}\widehat{n}(_\mu \widehat{n}\times _\nu \widehat{n})=_\mu \stackrel{~}{C}_\nu _\nu \stackrel{~}{C}_\mu ,`$ (8)
where $`\stackrel{~}{C}_\mu `$ is the โmagneticโ potential . This allows us to identify the non-Abelian monopole potential by
$`\stackrel{}{C}_\mu ={\displaystyle \frac{1}{g}}\widehat{n}\times _\mu \widehat{n},`$ (9)
in terms of which the magnetic field is expressed as
$`\stackrel{}{H}_{\mu \nu }=_\mu \stackrel{}{C}_\nu _\nu \stackrel{}{C}_\mu +g\stackrel{}{C}_\mu \times \stackrel{}{C}_\nu ={\displaystyle \frac{1}{g}}_\mu \widehat{n}\times _\nu \widehat{n}=H_{\mu \nu }\widehat{n}.`$ (10)
Notice that the magnetic field has a remarkable structure
$`H_{\mu \alpha }H_{\alpha \beta }H_{\beta \nu }={\displaystyle \frac{1}{2}}H_{\alpha \beta }^2H_{\mu \nu },`$ (11)
which will be very useful for us in the following.
With (1) one has
$`\stackrel{}{F}_{\mu \nu }=\widehat{F}_{\mu \nu }+\widehat{D}_\mu \stackrel{}{X}_\nu \widehat{D}_\nu \stackrel{}{X}_\mu +g\stackrel{}{X}_\mu \times \stackrel{}{X}_\nu ,`$ (12)
so that the Yang-Mills Lagrangian is expressed as
$`=`$ $``$ $`{\displaystyle \frac{1}{4}}\stackrel{}{F}_{\mu \nu }^2={\displaystyle \frac{1}{4}}\widehat{F}_{\mu \nu }^2{\displaystyle \frac{1}{4}}(\widehat{D}_\mu \stackrel{}{X}_\nu \widehat{D}_\nu \stackrel{}{X}_\mu )^2{\displaystyle \frac{g^2}{4}}(\stackrel{}{X}_\mu \times \stackrel{}{X}_\nu )^2`$ (13)
$``$ $`{\displaystyle \frac{g}{2}}\widehat{F}_{\mu \nu }(\stackrel{}{X}_\mu \times \stackrel{}{X}_\nu ).`$ (14)
This shows that the Yang-Mills theory can be viewed as the restricted gauge theory made of the dual potential $`\widehat{A}_\mu `$, which has the valence gluon $`\stackrel{}{X}_\mu `$ as its source . But notice that here the valence gluon has the magnetic moment interaction with the restricted potential. This interaction plays the crucial role in the monopole condensation as we will see in the following.
Obviously the theory is invariant under the gauge transformation (3) of the active type. But notice that it is also invariant under the following gauge transformation of the passive type,
$`\delta \widehat{n}=0,\delta \stackrel{}{A}_\mu ={\displaystyle \frac{1}{g}}D_\mu \stackrel{}{\alpha },`$ (15)
under which one has
$`\delta A_\mu ={\displaystyle \frac{1}{g}}\widehat{n}D_\mu \stackrel{}{\alpha },\delta \stackrel{}{C}_\mu =0,`$ (16)
$`\delta \stackrel{}{X}_\mu ={\displaystyle \frac{1}{g}}[D_\mu \stackrel{}{\alpha }(\widehat{n}D_\mu \stackrel{}{\alpha })\widehat{n}].`$ (17)
This gauge invariance of the passive type plays an important role in the background field method discussed in the following.
## III Dynamical Symmetry Breaking and Monopole Condensation
With this preparation we will now show that the effective action of QCD, which one obtains after integrating out all the dynamical degrees of the gluons from the monopole background, can be written in one loop approximation as
$`_g={\displaystyle \frac{Z}{4}}\stackrel{}{H}_{\mu \nu }^2,`$ (18)
$`Z=1+{\displaystyle \frac{22}{3}}{\displaystyle \frac{g^2}{(4\pi )^2}}\left(\mathrm{ln}{\displaystyle \frac{gH}{\mu ^2}}c\right),`$ (19)
where $`H=\sqrt{\stackrel{}{H}_{\mu \nu }^2}`$, $`\mu `$ is the modified minimal subtraction parameter, and $`c`$ is a constant. This generates the desired dynamical symmetry breaking and establishes the magnetic condensation of the vacuum.
To derive the effective action consider the generating functional of (10)
$`W[J_\mu ,\stackrel{}{J}_\mu ]`$ $`=`$ $`{\displaystyle }๐A_\mu ๐\stackrel{}{X}_\mu \mathrm{exp}\{i{\displaystyle }[{\displaystyle \frac{1}{4}}\widehat{F}_{\mu \nu }^2{\displaystyle \frac{1}{4}}(\widehat{D}_\mu \stackrel{}{X}_\nu \widehat{D}_\nu \stackrel{}{X}_\mu )^2`$ (20)
$``$ $`{\displaystyle \frac{g}{2}}\widehat{F}_{\mu \nu }(\stackrel{}{X}_\mu \times \stackrel{}{X}_\nu ){\displaystyle \frac{g^2}{4}}(\stackrel{}{X}_\mu \times \stackrel{}{X}_\nu )^2+A_\mu J_\mu +\stackrel{}{X}_\mu \stackrel{}{J}_\mu ]d^4x\}.`$ (21)
We have to perform the functional integral with a proper choice of a gauge, leaving $`\stackrel{}{C}_\mu `$ as a background. To do this we first fix the gauge with the condition
$`\stackrel{}{F}`$ $`=`$ $`\widehat{D}_\mu (A_\mu \widehat{n}+\stackrel{}{X}_\mu )=0,`$ (22)
$`_{gf}=`$ $``$ $`{\displaystyle \frac{1}{2\xi }}\left[(_\mu A_\mu )^2+(\widehat{D}_\mu \stackrel{}{X}_\mu )^2\right].`$ (23)
Notice that the gauge transformation of the passive type (11) plays the important role in this background field method. With the above gauge fixing the generating functional takes the following form,
$`W[J_\mu ,\stackrel{}{J}_\mu ]`$ $`=`$ $`{\displaystyle }๐A_\mu ๐\stackrel{}{X}_\mu ๐\stackrel{}{c}๐\stackrel{}{c}^{}\mathrm{exp}\{i{\displaystyle }[{\displaystyle \frac{1}{4}}\widehat{F}_{\mu \nu }^2`$ (24)
$``$ $`{\displaystyle \frac{1}{4}}(\widehat{D}_\mu \stackrel{}{X}_\nu \widehat{D}_\nu \stackrel{}{X}_\mu )^2{\displaystyle \frac{g}{2}}\widehat{F}_{\mu \nu }(\stackrel{}{X}_\mu \times \stackrel{}{X}_\nu ){\displaystyle \frac{g^2}{4}}(\stackrel{}{X}_\mu \times \stackrel{}{X}_\nu )^2`$ (25)
$`+`$ $`\stackrel{}{c}^{}\widehat{D}_\mu D_\mu \stackrel{}{c}{\displaystyle \frac{1}{2\xi }}(_\mu A_\mu )^2{\displaystyle \frac{1}{2\xi }}(\widehat{D}_\mu \stackrel{}{X}_\mu )^2`$ (26)
$`+`$ $`A_\mu J_\mu +\stackrel{}{X}_\mu \stackrel{}{J}_\mu ]d^4x\},`$ (27)
where $`\stackrel{}{c}`$ and $`\stackrel{}{c}^{}`$ are the ghost fields. In one loop approximation the $`A_\mu `$ integration becomes trivial, and the $`\stackrel{}{X}_\mu `$ and ghost integrations result in the following functional determinants (with $`\xi =1`$),
$`\mathrm{Det}^{\frac{1}{2}}K_{\mu \nu }^{ab}\mathrm{Det}^{\frac{1}{2}}[g_{\mu \nu }(\stackrel{~}{D}\stackrel{~}{D})^{ab}2gH_{\mu \nu }ฯต^{abc}n^c],`$ (28)
$`\mathrm{Det}M_{FP}^{ab}\mathrm{Det}[(\stackrel{~}{D}\stackrel{~}{D})^{ab}],`$ (29)
where $`\stackrel{~}{D}_\mu `$ is defined with only the background $`\stackrel{}{C}_\mu `$. One can simplify the determinant $`K`$ using the relation (8),
$`\mathrm{ln}\mathrm{Det}^{\frac{1}{2}}K`$ $`=`$ $`\mathrm{ln}\mathrm{Det}[(\stackrel{~}{D}\stackrel{~}{D})^{ab}]`$ (30)
$``$ $`{\displaystyle \frac{1}{2}}\mathrm{ln}\mathrm{Det}[(\stackrel{~}{D}\stackrel{~}{D})^{ab}+i\sqrt{2}gHฯต^{abc}n^c]`$ (31)
$``$ $`{\displaystyle \frac{1}{2}}\mathrm{ln}\mathrm{Det}[(\stackrel{~}{D}\stackrel{~}{D})^{ab}i\sqrt{2}gHฯต^{abc}n^c].`$ (32)
With this the one loop contribution of the functional determinants to the effective action can be written as
$`\mathrm{\Delta }S=i\mathrm{ln}\mathrm{Det}[(\stackrel{~}{D}^2+\sqrt{2}gH)(\stackrel{~}{D}^2\sqrt{2}gH)],`$ (33)
where now $`\stackrel{~}{D}_\mu `$ acquires the following Abelian form,
$`\stackrel{~}{D}_\mu =_\mu +ig\stackrel{~}{C}_\mu .`$ (34)
Notice that the reason for this simplification is precisely because our restricted potential $`\widehat{A}_\mu `$ originates from the Abelian projection.
With this one can use the heat kernel method and calculate the functional determinant. For a covariantly constant $`\stackrel{}{H}_{\mu \nu }`$ one finds
$`\mathrm{\Delta }={\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{2ฯต}}}{\displaystyle \frac{gH/\sqrt{2}\mu ^2}{\mathrm{sinh}(gHt/\sqrt{2}\mu ^2)}}`$ (35)
$`\times [\mathrm{exp}(\sqrt{2}gHt/\mu ^2)+\mathrm{exp}(\sqrt{2}gHt/\mu ^2)],`$ (36)
where $`ฯต`$ is the ultra-violet cut-off parameter. The integral contains the (usual) ultra-violet divergence, but notice that here it is also plagued by a severe infra-red divergence. This, of course, is precisely what one should have expected, because such an infra-red divergence is an unavoidable characteristics of QCD. So the important issue now becomes how to regularize the infra-red divergence.
To find the correct infra-red regularization, one must understand the origin of the divergence. The infra-red divergence can be traced back to the magnetic moment interaction of the gluons that we have in (10), which is also well-known to be responsible for the asymptotic freedom . This magnetic interaction generates negative eigenvalues in Det K in the long distance region, which cause the infra-red divergence. More precisely when the momentum $`k`$ of the gluon parallel to the background magnetic field becomes smaller than the background field strength (i.e., when $`k^2<gH/\sqrt{2}`$), the lowest Landau level gluon eigenfunction whose spin is parallel to the magnetic field acquires an imaginary energy and thus becomes tachyonic. It is these unphysical tachyonic states which cause the infra-red divergence. So one must exclude these tachyonic modes in the calculation of the effective action, when one makes a proper infra-red regularization. Including the tachyons in the physical spectrum will surely destablize QCD and make it ill-defined.
With this understanding we can do both the ultra-violet and infra-red regularizations simultaneously with due care. Excluding the contribution of the unphysical modes we have
$`\mathrm{\Delta }={\displaystyle \frac{11g^2}{96\pi ^2}}H^2({\displaystyle \frac{1}{ฯต}}\gamma ){\displaystyle \frac{11g^2}{96\pi ^2}}H^2(\mathrm{ln}{\displaystyle \frac{gH}{\mu ^2}}c_1)`$ (37)
$`c_1=1{\displaystyle \frac{1}{2}}\mathrm{ln}2{\displaystyle \frac{24}{11}}\zeta ^{}(1,{\displaystyle \frac{3}{2}})=1.29214\mathrm{},`$ (38)
where $`\gamma `$ is the Eulerโs constant and $`\zeta ^{}(x,y)`$ is the generalized Hurwitz zeta-function. From this we finally obtain (with the modified minimal subtraction)
$`_g={\displaystyle \frac{1}{4}}H^2{\displaystyle \frac{11g^2}{96\pi ^2}}H^2(\mathrm{ln}{\displaystyle \frac{gH}{\mu ^2}}c_1).`$ (39)
This completes the derivation of the effective action (13).
Clearly the effective action provides the following non-trivial effective potential
$`V={\displaystyle \frac{g^2}{4}}(\stackrel{}{C}_\mu \times \stackrel{}{C}_\nu )^2\left\{1+{\displaystyle \frac{22}{3}}{\displaystyle \frac{g^2}{(4\pi )^2}}\left[\mathrm{ln}{\displaystyle \frac{g^2[(\stackrel{}{C}_\mu \times \stackrel{}{C}_\nu )^2]^{1/2}}{\mu ^2}}c_1\right]\right\},`$ (40)
which generates the desired monopole condensation of the vacuum,
$`<H>={\displaystyle \frac{\mu ^2}{g}}\mathrm{exp}\left({\displaystyle \frac{24\pi ^2}{11g^2}}{\displaystyle \frac{1}{2}}+c_1\right).`$ (41)
Notice that with $`\alpha _s=1`$ we have
$`{\displaystyle \frac{<H>}{\mu ^2}}=0.11225\mathrm{}.`$ (42)
The vacuum generates an โeffective massโ for $`\stackrel{}{C}_\mu `$,
$`m^2={\displaystyle \frac{11g^4}{96\pi ^2}}<{\displaystyle \frac{(\stackrel{}{C}_\mu \times \stackrel{}{H}_{\mu \nu })^2}{H^2}}>,`$ (43)
which demonstrates that the monopole condensation indeed generates the mass gap necessary for the dual Meissner effect. Obviously the mass scale sets the confinement scale.
To check the consistency of our result with the perturbative QCD we now discuss the running coupling and the renormalization. For this we define the running coupling $`\overline{g}`$ by
$`{\displaystyle \frac{^2V}{H^2}}|_{H=\overline{H}}={\displaystyle \frac{1}{2}}{\displaystyle \frac{g^2}{\overline{g}^2}}.`$ (44)
So with $`g\overline{H}=\overline{\mu }^2\mathrm{exp}(c_13/2)`$ we obtain the following $`\beta `$-function,
$`{\displaystyle \frac{1}{\overline{g}^2}}={\displaystyle \frac{1}{g^2}}+{\displaystyle \frac{11}{12\pi ^2}}\mathrm{ln}{\displaystyle \frac{\overline{\mu }}{\mu }},\beta (\overline{\mu })={\displaystyle \frac{11}{24\pi ^2}}\overline{g}^3,`$ (45)
which exactly coincides with the well-known asymptotic freedom result . This confirms that the asymptotic freedom and the monopole condensation have exactly the same underlying dynamics.
In terms of the running coupling the renormalized potential is given by
$`V_{\mathrm{ren}}={\displaystyle \frac{1}{4}}H^2\left[1+{\displaystyle \frac{22}{3}}{\displaystyle \frac{\overline{g}^2}{(4\pi )^2}}(\mathrm{ln}{\displaystyle \frac{\overline{g}H}{\overline{\mu }^2}}c_1)\right],`$ (46)
and the Callan-Symanzik equation
$`\left(\overline{\mu }{\displaystyle \frac{}{\overline{\mu }}}+\beta {\displaystyle \frac{}{\overline{g}}}\gamma (\stackrel{}{C}_\mu )\stackrel{}{C}_\mu {\displaystyle \frac{}{\stackrel{}{C}_\mu }}\right)V_{\mathrm{ren}}=0`$ (47)
gives the following anomalous dimension for $`\stackrel{}{C}_\mu `$,
$`\gamma (\stackrel{}{C}_\mu )={\displaystyle \frac{11}{24\pi ^2}}\overline{g}^2.`$ (48)
This should be compared with that of the gluon field in perturbative QCD, $`\gamma (\stackrel{}{A}_\mu )=5\overline{g}^2/24\pi ^2`$ for $`SU(2)`$.
There have been many attempts to construct the effective action of QCD in the literature, and in the appearance our vacuum (24) looks very much like the old Savvidy-Nielsen-Olesen vacuum . But it must be emphasized that there are fundamental differences between the earlier attempts and the present approach. The earlier attempts had two problems. First the separation between the classical background and the quantum field was not gauge independent, which made the one loop vacuum neither gauge invariant nor Lorentz invariant. This violation of the gauge invariance was of course a serious defect, but perhaps the more serious problem was that the infra-red divergence was not properly regularized in many of the earlier attempts. Indeed it has been asserted that the Savvidy-Nielsen-Olesen vacuum should be unstable, because the effective action which defines the vacuum develops an imaginary part ,
$`Im_g|_{SNO}={\displaystyle \frac{g^2}{16\pi }}H^2,`$ (49)
which destabilizes the vacuum through the pair creation of gluons. This assertion of the instability of the Savvidy-Nielsen-Olesen vacuum, which comes from improper infra-red regularizations, has been widely accepted and never been convincingly revoked. Obviously, without a proper infra-red regularization one can not expect to obtain the correct effective action of QCD. Because of these defects the earlier attempts have not been so successful.
In contrast in our approach the separation of the monopole background from the quantum fluctuation is clearly gauge independent. Moreover our infra-red regularization generates no imaginary part in the effective action. Because of these we obtain a stable vacuum made of monopole condensation which is both gauge and Lorentz invariant. Notice that the infra-red regularization in (20) is not just to remove the infra-red divergence (there are infinitely many ways to do this). The infra-red divergence that we face here in QCD is also different from those one encounters in the massless QED. The infra-red divergence in the massless QED comes from the zero modes. But these zero modes are physical modes, which should not be excluded in the calculation of the effective action. On the other hand the infra-red divergence that we have here comes from the unphysical modes, and one must exclude these unphysical modes from the physical spectrum with a proper infra-red regularization (Notice that in the earlier attempts these tachyonic modes are incorrectly identified as the โunstableโ modes, but we emphasize that they are not just unstable but unphysical). And it is precisely these unphysical modes that generate the controversial imaginary part in the Savvidy-Nielsen-Olesen action. So with the exclusion of the unphysical modes the instability of the vacuum disappears completely.
As importantly in our approach we can really claim that the magnetic condensation is a gauge independent phenomenon. Furthermore here we have demonstrated that it is precisely the Wu-Yang monopole that is responsible for the condensation. Notice that in the earlier attempts it has never been clear what was the source of the condensation, nor has it been simple to show that the condensation is indeed a gauge independent phenomenon.
Clearly the quark loop makes an additional contribution to the effective action. As we will see in the following we have for the massless quarks,
$`\mathrm{\Delta }_q={\displaystyle \frac{g^2H^2}{96\pi ^2}}N_f(\mathrm{ln}{\displaystyle \frac{gH}{\mu ^2}}c_2),`$ (50)
$`c_2=\gamma +{\displaystyle \frac{1}{2}}\mathrm{ln}2+\mathrm{ln}(2\pi )+{\displaystyle \frac{6}{\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n^2}}\mathrm{ln}n=3.33163\mathrm{},`$ (51)
where $`N_f`$ is the number of the flavors of the quarks. With this we obtain the following total effective action of $`SU(2)`$ QCD for the pure magnetic background,
$`_{eff}`$ $`=`$ $`{\displaystyle \frac{1}{4}}H^2{\displaystyle \frac{11N_f}{96\pi ^2}}g^2H^2(\mathrm{ln}{\displaystyle \frac{gH}{\mu ^2}}c_3),`$ (52)
$`c_3`$ $`={\displaystyle \frac{11}{11N_f}}c_1{\displaystyle \frac{N_f}{11N_f}}c_2.`$ (53)
The corresponding effective potential is plotted in Fig.1, where we have assumed $`\alpha _s=1,\mu =1`$, and $`N_f=2`$. The effective potential clearly shows that there is indeed a dynamical symmetry breaking in QCD.
## IV Electric Background
To make sure that our infra-red regularization is indeed the correct one it is necessary to have an independent confirmation of the above result. To do this it is instructive to calculate the effective action with a pure electric background first.
So let us consider the general case where the background field strengh $`\widehat{F}_{\mu \nu }`$ contains both electric and magnetic components. In this case we have
$`\widehat{F}_{\mu \nu }=G_{\mu \nu }\widehat{n},G_{\mu \nu }=F_{\mu \nu }+H_{\mu \nu },`$ (54)
and the functional determinants of the gluon and the ghost loops are generalized to
$`\mathrm{Det}^{\frac{1}{2}}K_{\mu \nu }^{ab}\mathrm{Det}^{\frac{1}{2}}[g_{\mu \nu }(\widehat{D}\widehat{D})^{ab}2gG_{\mu \nu }ฯต^{abc}n^c],`$ (55)
$`\mathrm{Det}M_{FP}=\mathrm{Det}[(\widehat{D}\widehat{D})^{ab}],`$ (56)
where now $`\widehat{D}_\mu `$ is defined with an arbitrary background field $`\widehat{A}_\mu `$. Using the relation
$`G_{\mu \alpha }G_{\alpha \beta }G_{\beta \nu }={\displaystyle \frac{1}{2}}G^2G_{\mu \nu }{\displaystyle \frac{1}{2}}(G\stackrel{~}{G})\stackrel{~}{G}_{\mu \nu }(\stackrel{~}{G}_{\mu \nu }={\displaystyle \frac{1}{2}}ฯต_{\mu \nu \rho \sigma }G_{\rho \sigma }),`$ (57)
one can simplify the functional determinants of the gluon and the quark loops as follows,
$`\mathrm{ln}\mathrm{Det}^{\frac{1}{2}}[(g_{\mu \nu }(\widehat{D}\widehat{D})^{ab}2gG_{\mu \nu }ฯต^{abc}n^c]=`$ (58)
$`\mathrm{ln}\mathrm{Det}[(\stackrel{~}{D}^2+2a)(\stackrel{~}{D}^22a)(\stackrel{~}{D}^22ib)(\stackrel{~}{D}^2+2ib)],`$ (59)
$`\mathrm{ln}\mathrm{Det}M_{FP}=2\mathrm{ln}\mathrm{Det}(\stackrel{~}{D}^2),`$ (60)
where
$`a={\displaystyle \frac{g}{2}}\sqrt{\sqrt{G^4+(G\stackrel{~}{G})^2}+G^2},b={\displaystyle \frac{g}{2}}\sqrt{\sqrt{G^4+(G\stackrel{~}{G})^2}G^2},`$ (61)
and now $`\stackrel{~}{D}_\mu `$ is defined with an arbitrary background $`A_\mu +\stackrel{~}{C}_\mu `$,
$`\stackrel{~}{D}_\mu =+ig(A_\mu +\stackrel{~}{C}_\mu ).`$ (62)
So for a pure electric background (i.e., for $`a=0`$) we have
$`\mathrm{\Delta }={\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{2ฯต}}}{\displaystyle \frac{b}{\mathrm{sin}(bt)}}[\mathrm{exp}(2ibt)+\mathrm{exp}(2ibt)].`$ (63)
Notice that (unlike the pure magnetic background) the integrand of the above integral has poles on the real axis, so that we must specify the contour of the integral. Here the causality requires the contour to pass above the real axis.
There are different ways to evaluate the integral, but a simple and nice way of doing this follows from the observation that in the imaginary time (i.e., in the Minkowski time) the role of the electric and magnetic fields are reversed. So with the Wick rotation of the proper time $`t`$ to the imaginary time $`it`$, the above integral acquires the same form as (20). Indeed with the Wick rotation (39) becomes
$`\mathrm{\Delta }{\displaystyle \frac{1}{16\pi ^2}}i^ฯต{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{2ฯต}}}{\displaystyle \frac{b}{\mathrm{sinh}(bt)}}[\mathrm{exp}(2bt)+\mathrm{exp}(2bt)].`$ (64)
From this we obtain
$`\mathrm{\Delta }={\displaystyle \frac{11b^2}{48\pi ^2}}({\displaystyle \frac{1}{ฯต}}\gamma )+{\displaystyle \frac{11b^2}{48\pi ^2}}(\mathrm{ln}{\displaystyle \frac{b}{\mu ^2}}c_g)i{\displaystyle \frac{11b^2}{96\pi }},`$ (65)
$`c_g=1\mathrm{ln}2{\displaystyle \frac{24}{11}}\zeta ^{}(1,{\displaystyle \frac{3}{2}})=0.94556\mathrm{}.`$ (66)
So with the modified minimal subtraction we have (with the pure electric background)
$`_g={\displaystyle \frac{b^2}{2g^2}}+{\displaystyle \frac{11b^2}{48\pi ^2}}(\mathrm{ln}{\displaystyle \frac{b}{\mu ^2}}c_g)i{\displaystyle \frac{11b^2}{96\pi }}.`$ (67)
It must be emphasized that in evaluating the above integral the same infra-red regularization is applied as in the pure magnetic background. With the pure electric background the eigenfunctions of Det K in the long distance region (i.e., for $`k^2<b`$) become anti-causal and thus unphysical, just like the eigenfunctions under the pure magnetic background become tachyonic and unphysical in the infra-red region (i.e., for $`k^2<gH/\sqrt{2})`$. So we must again exclude these unphysical modes to evaluate the above integral.
The contrast between the effective actions (22) and (42) is remarkable. First, (42) has no local minimum. This implies that the electric background does not generate a condensation. Secondly, (42) has an imaginary part
$`Im_g={\displaystyle \frac{11b^2}{96\pi }}.`$ (68)
This implies that the electric background is unstable. But perhaps a more important point here is that the imaginary part is negative. This means that the electric background generates the pair annihilation, rather than the pair creation, of the gluons. This is because the negative imaginary part can be interpreted as the negative probability of the pair creation. This implies that the gluons in QCD, unlike the electrons in QED, tend to annihilate among themselves in the color electric field. This is really remarkable because this is precisely what one needs to explain the asymptotic freedom. Remember that the asymptotic freedom comes from the anti-screening effect, but for this anti-screening one needs the pair annihilation of gluons in the color electric flux. This means that our result is not only consistent with the asymptotic freedom, but actually explains why one must have the asymptotic freedom in QCD.
With this we can now make an independent confirmation of our effective actions (22) and (42). To do this first notice that the imaginary part (32) of the Savvidy-Nielsen-Olesen action as well as ours (22) and (42) are quadratic in the background fields. In our notation (38) this means that the imaginary part of the one loop effective action is second order in the coupling constant $`g`$. So one can find the correct imaginary part of the effective action perturbatively, just by calculating the effective action up to the second order in the coupling constant in the perturbative expansion. Now, a remarkable point is that by doing this one can reproduce our results ,
$`Im\mathrm{\Delta }_g=\{{\displaystyle \genfrac{}{}{0pt}{}{0b=0}{{\displaystyle \frac{11b^2}{96\pi }}a=0.}}`$ (69)
This confirms that our infra-red regularization is indeed correct. More importantly this confirms that we do have the desired dynamical symmetry breaking and the magnetic condensation in QCD. In the following we will provide a third independent argument which supports our results.
An important point to observe here is that the effective actions (22) and (42) are actually the mirror image of each other. To see this notice that we can obtain (42) from (22) simply by replacing $`a=gH/\sqrt{2}`$ with $`ib`$, and similarly (22) from (42) by replacing $`b`$ with $`ia=igH/\sqrt{2}`$. This is the first indication that there exists a fundamental symmetry which we call the duality, in the effective action of QCD. We will discuss this duality in detail in the following.
The quark loop makes an extra contribution to the effective action. As we will see in the following we find for the massless quarks,
$`\mathrm{\Delta }_q`$ $`=`$ $`{\displaystyle \frac{b^2}{48\pi ^2}}N_f(\mathrm{ln}{\displaystyle \frac{b}{\mu ^2}}c_q)+i{\displaystyle \frac{b^2}{96\pi }},`$ (70)
$`c_q`$ $`=`$ $`\gamma +\mathrm{ln}(2\pi )+{\displaystyle \frac{6}{\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n^2}}\mathrm{ln}n.`$ (71)
So, together with (42), we obtain the following effective action
$`_{eff}={\displaystyle \frac{b^2}{2g^2}}+{\displaystyle \frac{11N_f}{48\pi ^2}}b^2(\mathrm{ln}{\displaystyle \frac{b}{\mu ^2}}c_t),`$ (72)
$`c_t={\displaystyle \frac{11}{11N_f}}c_g{\displaystyle \frac{1}{11N_f}}c_q,`$ (73)
which is shown in Fig.2.
## V Vacuum Stability
So far we have established the monopole condensation as a dynamical symmetry breaking. But this by itself does not allow us to claim that the monopole condensation provides the true physical vacuum. To show that the monopole condensation is indeed the unique vacuum of QCD, one must calculate the effective action with an arbitrary background of the restricted potential $`\widehat{A}_\mu `$ and show that indeed the monopole condensation provides the true stable minimum of the effective potential.
For a general background with arbitrary $`a`$ and $`b`$, the contribution of the gluon and ghost loops corresponding to the functional determinant (37) is given by
$`\mathrm{\Delta }_g={\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{3ฯต}}}{\displaystyle \frac{abt^2}{\mathrm{sinh}(at)\mathrm{sin}(bt)}}`$ (74)
$`\times [\mathrm{exp}(2at)+\mathrm{exp}(2at)+\mathrm{exp}(2ibt)+\mathrm{exp}(2ibt)2]`$ (75)
$`=\mathrm{\Delta }_1+\mathrm{\Delta }_2+\mathrm{\Delta }_3,`$ (76)
where
$`\mathrm{\Delta }_1={\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{3ฯต}}}{\displaystyle \frac{abt^2}{\mathrm{sinh}(at)\mathrm{sin}(bt)}}[\mathrm{exp}(2at)+\mathrm{exp}(2at)],`$ (77)
$`\mathrm{\Delta }_2={\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{3ฯต}}}{\displaystyle \frac{abt^2}{\mathrm{sinh}(at)\mathrm{sin}(bt)}}[\mathrm{exp}(2ibt)+\mathrm{exp}(2ibt)],`$ (78)
$`\mathrm{\Delta }_3={\displaystyle \frac{1}{8\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{3ฯต}}}{\displaystyle \frac{abt^2}{\mathrm{sinh}(at)\mathrm{sin}(bt)}}.`$ (79)
Notice that $`\mathrm{\Delta }_1+\mathrm{\Delta }_2`$ describes the contribution of Det K of the gluon loop, but $`\mathrm{\Delta }_3`$ describes the contribution of Det M of the ghost loop. Here again one should keep in mind that the contour of the integral should pass above the $`t`$-axis to preserve the causality.
The integral expression (47) of the effective action has been known for some time , but the actual integration of it is not easy to perform. Indeed, as far as we understand, the integration has not been completed satisfactorily. In the following we will perform the integral, and present a compact expression of the effective action. To carry out the integral we need to re-express the integrand in such a way that we can do the integral analytically. For this purpose we introduce the following identity ,
$`{\displaystyle \frac{xy}{\mathrm{sinh}(x)\mathrm{sin}(y)}}`$ $`=`$ $`1{\displaystyle \frac{x^2y^2}{6}}{\displaystyle \frac{2}{\pi }}x^3y{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}{\displaystyle \frac{\mathrm{csch}({\displaystyle \frac{n\pi y}{x}})}{x^2+n^2\pi ^2}}`$ (80)
$`+`$ $`{\displaystyle \frac{2}{\pi }}xy^3{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}{\displaystyle \frac{\mathrm{csch}({\displaystyle \frac{n\pi x}{y}})}{y^2n^2\pi ^2}}.`$ (81)
The identity, which we can establish using one of the Ramanujianโs identities, has played an important role in the calculation of the effective action of the scalar QED. Here again in QCD the same identity plays the crucial role in evaluating the effective action.
Now we can calculate $`\mathrm{\Delta }_1,\mathrm{\Delta }_2`$, and $`\mathrm{\Delta }_3`$ separately with the help of our identity. Indeed using the identity (49) we obtain the following expression for $`\mathrm{\Delta }_1`$,
$`\mathrm{\Delta }_1=I_1(0,2a)+I_1(0,2a)+I_2(0,2a)+I_2(0,2a)+I_3(0,2a)+I_3(0,2a),`$ (82)
where
$`I_1(ฯต,\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}t^{ฯต3}\left(1{\displaystyle \frac{a^2b^2}{6}}t^2\right)e^{\lambda t}๐t`$ (83)
$`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}\lambda ^ฯต\left[{\displaystyle \frac{\lambda ^2}{2}}(1+{\displaystyle \frac{3}{2}}ฯต){\displaystyle \frac{a^2b^2}{6}}\right]\mathrm{\Gamma }(ฯต),`$ (84)
$`I_2(ฯต,\lambda )`$ $`=`$ $`{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi b}{a}}){\displaystyle _0^{\mathrm{}}}t^{ฯต+1}{\displaystyle \frac{e^{\lambda t}}{t^2+({\displaystyle \frac{n\pi }{a}})^2}}๐t`$ (85)
$`=`$ $`{\displaystyle \frac{ab}{8\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})[\mathrm{ci}({\displaystyle \frac{n\pi \lambda }{a}})\mathrm{cos}({\displaystyle \frac{n\pi \lambda }{a}})+\mathrm{si}({\displaystyle \frac{n\pi \lambda }{a}})\mathrm{sin}({\displaystyle \frac{n\pi \lambda }{a}})],`$ (86)
$`I_3(ฯต,\lambda )`$ $`=`$ $`{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}}){\displaystyle _0^{\mathrm{}}}t^{ฯต+1}{\displaystyle \frac{e^{\lambda t}}{t^2({\displaystyle \frac{n\pi }{b}})^2}}๐t`$ (87)
$`=`$ $`{\displaystyle \frac{ab}{16\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})[\mathrm{Ei}({\displaystyle \frac{n\pi \lambda }{b}})\mathrm{exp}({\displaystyle \frac{n\pi \lambda }{a}})+\mathrm{Ei}({\displaystyle \frac{n\pi \lambda }{b}})\mathrm{exp}({\displaystyle \frac{n\pi \lambda }{a}})].`$ (88)
Notice that here ci($`x`$) and si($`x`$) are the cosine and sine integral functions, and Ei($`x`$) is the exponential integral function ,
$`\mathrm{ci}(x)`$ $`=`$ $`{\displaystyle _x^{\mathrm{}}}{\displaystyle \frac{\mathrm{cos}(t)}{t}}๐t`$ (89)
$`=`$ $`\gamma +\mathrm{ln}x+{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^nx^{2n}}{(2n)!2n}},(\mathrm{Re}x>0)`$ (90)
$`\mathrm{si}(x)`$ $`=`$ $`{\displaystyle _x^{\mathrm{}}}{\displaystyle \frac{\mathrm{sin}(t)}{t}}๐t`$ (91)
$`=`$ $`{\displaystyle \frac{\pi }{2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^nx^{2n1}}{(2n1)!(2n1)}},(\mathrm{Re}x>0)`$ (92)
$`\mathrm{Ei}(x)`$ $`=`$ $`{\displaystyle _x^{\mathrm{}}}{\displaystyle \frac{e^t}{t}}๐t`$ (93)
$`=`$ $`\gamma +\mathrm{ln}x+{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^nx^n}{(n)!n}}.(\mathrm{Re}x>0)`$ (94)
With these we find
$`\mathrm{\Delta }_1`$ $`=`$ $`{\displaystyle \frac{11a^2+b^2}{48\pi ^2}}({\displaystyle \frac{1}{ฯต}}\gamma )+{\displaystyle \frac{3a^2}{8\pi ^2}}{\displaystyle \frac{11a^2+b^2}{48\pi ^2}}\mathrm{ln}({\displaystyle \frac{2a}{\mu ^2}})`$ (95)
$`+`$ $`{\displaystyle \frac{ab}{4\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{ci}(2n\pi )`$ (96)
$``$ $`{\displaystyle \frac{ab}{8\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})[\mathrm{Ei}({\displaystyle \frac{2n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}})`$ (97)
$`+`$ $`\overline{\mathrm{Ei}}({\displaystyle \frac{2n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}})]+i{\displaystyle \frac{a^2}{96\pi }}`$ (98)
$``$ $`i{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\left[\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})+\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}})\right],`$ (99)
where $`\overline{\mathrm{Ei}}(x)`$ is $`\mathrm{Re}\mathrm{Ei}(x)`$. To obtain the above result notice that, for $`\lambda =2a`$, a naive integration of (51) gives an infra-red divergence. So one must perform the integral with a positive $`\lambda `$ first, and then make an analytic continuation to $`\lambda =2a`$. In this analytic continuation one must keep in mind two points. First, the analytic continuation should preserve the causality. This means that we must select the correct (i.e., physical) branch in the analytic continuation of ci($`x`$), si($`x`$), and Ei($`x`$) to the negative real axis. Secondly, the continuation must be done in such a way that the unphysical modes should have no contribution. This requires that the imaginary component of $`\mathrm{\Delta }_1`$ should reproduce the previous result (21) in the pure magnetic background. With these precautions we obtain the above result.
Now it is simple to evaluate $`\mathrm{\Delta }_2`$, because it can be put into the same form as $`\mathrm{\Delta }_1`$ with a Wick rotation of $`t`$ to $`it`$. Indeed we have (after the Wick rotation)
$`\mathrm{\Delta }_2={\displaystyle \frac{1}{16\pi ^2}}i^ฯต{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{3ฯต}}}{\displaystyle \frac{abt^2}{\mathrm{sinh}(bt)\mathrm{sin}(at)}}[\mathrm{exp}(2bt)+\mathrm{exp}(2bt)].`$ (100)
But an important point to notice here is that with the Wick rotation the contour of the above integral should now pass below the $`t`$-axis. With this observation we obtain
$`\mathrm{\Delta }_2=`$ $``$ $`{\displaystyle \frac{11b^2+a^2}{48\pi ^2}}({\displaystyle \frac{1}{ฯต}}\gamma ){\displaystyle \frac{3b^2}{8\pi ^2}}+{\displaystyle \frac{11b^2+a^2}{48\pi ^2}}\mathrm{ln}({\displaystyle \frac{2b}{\mu ^2}})`$ (101)
$``$ $`{\displaystyle \frac{ab}{4\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{ci}(2n\pi )`$ (102)
$`+`$ $`{\displaystyle \frac{ab}{8\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})[\mathrm{Ei}({\displaystyle \frac{2n\pi b}{a}})\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})`$ (103)
$`+`$ $`\overline{\mathrm{Ei}}({\displaystyle \frac{2n\pi b}{a}})\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})]i{\displaystyle \frac{11b^2}{96\pi }}`$ (104)
$`+`$ $`i{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\left[\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})\right].`$ (105)
Again we emphasize that the above Wick rotation prescription automatically and naturally guarantees that we have the same infra-red regularization in the evaluation of $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$. Indeed we can easily confirm that $`\mathrm{\Delta }_2`$ reproduces the previous result (41) in the pure electric background (i.e., in the limit $`a`$ goes to zero), as it should.
Finally it is straightforward to calculate $`\mathrm{\Delta }_3`$, because it has no infra-red divergence. We find
$`\mathrm{\Delta }_3`$ $`=`$ $`{\displaystyle \frac{1}{8\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{3ฯต}}}{\displaystyle \frac{abt^2}{\mathrm{sinh}(at)\mathrm{sin}(bt)}}`$ (106)
$`=`$ $`2\left(I_1(0,0)+I_2(0,0)+I_3(0,0)\right).`$ (107)
From this we have
$`\mathrm{\Delta }_3={\displaystyle \frac{a^2b^2}{48\pi ^2}}({\displaystyle \frac{1}{ฯต}}\gamma ){\displaystyle \frac{ab}{4\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}[\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})(\mathrm{ln}({\displaystyle \frac{n\pi \mu ^2}{a}})+\gamma )`$ (108)
$`\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})(\mathrm{ln}({\displaystyle \frac{n\pi \mu ^2}{b}})+\gamma )]+i{\displaystyle \frac{ab}{8\pi ^3}}{\displaystyle }_{n=1}^{\mathrm{}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}}).`$ (109)
To obtain the above result we have used the following identity ,
$`{\displaystyle \frac{12}{\pi }}ab{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\left[\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\right]=b^2a^2.`$ (110)
In evaluating the integrals it should be stressed again that one must be very careful to implement a proper infra-red regularization. In particular, one has to make sure that the above result reproduces the results of the previous sections for the pure magnetic and the pure electric backgrounds.
With the above results we finally obtain (after the modified minimal subtraction)
$`\mathrm{\Delta }_g`$ $`=`$ $`\mathrm{\Delta }_1+\mathrm{\Delta }_2+\mathrm{\Delta }_3`$ (111)
$`=`$ $`{\displaystyle \frac{3a^23b^2}{8\pi ^3}}{\displaystyle \frac{11a^2+b^2}{48\pi ^2}}\mathrm{ln}({\displaystyle \frac{2a}{\mu ^2}})+{\displaystyle \frac{11b^2+a^2}{48\pi ^2}}\mathrm{ln}({\displaystyle \frac{2b}{\mu ^2}})`$ (112)
$`+`$ $`{\displaystyle \frac{ab}{4\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\left(\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\right)(\mathrm{ci}(2n\pi )\gamma )`$ (113)
$`+`$ $`{\displaystyle \frac{ab}{8\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}[\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})(\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})\mathrm{Ei}({\displaystyle \frac{2n\pi b}{a}})`$ (114)
$`+`$ $`\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})\overline{\mathrm{Ei}}({\displaystyle \frac{2n\pi b}{a}}))\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})(\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}})\mathrm{Ei}({\displaystyle \frac{2n\pi a}{b}})`$ (115)
$`+`$ $`\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}})\overline{\mathrm{Ei}}({\displaystyle \frac{2n\pi a}{b}}))]`$ (116)
$``$ $`{\displaystyle \frac{ab}{4\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\left(\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{ln}({\displaystyle \frac{n\pi \mu ^2}{a}})\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{ln}({\displaystyle \frac{n\pi \mu ^2}{b}})\right)`$ (117)
$``$ $`i{\displaystyle \frac{a^2+11b^2}{96\pi }}i{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}(\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})`$ (118)
$`+`$ $`\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}}))+i{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle }_{n=1}^{\mathrm{}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}}).`$ (119)
Notice again that here we have used our identity (58) to obtain the above result. The result is summarized in Fig.3 and Fig.4, where we have plotted the gluon contribution of the dispersive and absorbtive parts of the effective action.
An important point here is that the effective action acquires the following imaginary component,
$`Im\mathrm{\Delta }_g=`$ $``$ $`{\displaystyle \frac{a^2+11b^2}{96\pi }}{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}(\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})`$ (120)
$`+`$ $`\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}}))+{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle }_{n=1}^{\mathrm{}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}}).`$ (121)
We can confirm that this expression reproduces the previous results. Indeed we find
$`Im\mathrm{\Delta }_1=\{{\displaystyle \genfrac{}{}{0pt}{}{0b=0}{{\displaystyle \frac{b^2}{96\pi }}a=0,}}`$ (122)
$`Im\mathrm{\Delta }_2=\{{\displaystyle \genfrac{}{}{0pt}{}{0b=0}{{\displaystyle \frac{11b^2}{96\pi }}a=0,}}`$ (123)
$`Im\mathrm{\Delta }_3=\{{\displaystyle \genfrac{}{}{0pt}{}{0b=0}{{\displaystyle \frac{b^2}{96\pi }}a=0,}}`$ (124)
so that
$`Im\mathrm{\Delta }_g=\{{\displaystyle \genfrac{}{}{0pt}{}{0b=0}{{\displaystyle \frac{11b^2}{96\pi }}a=0.}}`$ (125)
What is really remarkable about the above result is the unilateral emergence of the imaginary part with $`b0`$. This immediately tells that the color electric background generates an instability to the effective action. Only when the background becomes pure magnetic the imaginary part disappears completely. This automatically proves that the monopole condensation is indeed the unique and stable vacuum of QCD, at least in one loop approximation.
Observe that the imaginary part of the effective action becomes negative in general. This assures again that the electric background makes the pair annihilation for the gluons, but not the pair creation. In contrast, as we will see in the following, the electric background makes the pair creation for the quarks. This is a very important observation, because this tells that the gluon pairs behave differently from the quark pairs in the color electric field. In particular the valence gluons are not likely to form the glueball bound states. This explains the experimental fact that there are so few (if at all) candidates of glueball bound states, while we have towers of hadronic bound states made of quarks.
## VI Quark Contribution
Obviously one can not neglect the quarks in QCD. Let us consider the Lagrangian involving the $`SU(2)`$ quarks in the fundamental representation
$`_q=\overline{\mathrm{\Psi }}(i\gamma ^\mu D_\mu m_q)\mathrm{\Psi },`$ (126)
$`D_\mu =_\mu +{\displaystyle \frac{g}{2i}}\stackrel{}{\sigma }\stackrel{}{A}_\mu ,`$ (127)
where $`m_q`$ is the mass of the quarks. One can express the quark contribution to the effective action in one loop approximation by
$`\mathrm{\Delta }_q={\displaystyle \frac{ab}{16\pi ^2}}N_f{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t^{1ฯต}}}\mathrm{coth}(at)\mathrm{cot}(bt)\mathrm{exp}(m_q^2t),`$ (128)
Notice that formally this is very much like the well-known expression of the one loop contribution of electron to the effective action in QED . This is because at one loop level only the interaction of the quarks with the restricted potential contributes to the effective action.
We can evaluate the above integral exactly the same way as we calculate the electron loop contribution in QED . Using the following Sitaramachandraraoโs identity ,
$`xy\mathrm{coth}(x)\mathrm{cot}(y)`$ $`=`$ $`1+{\displaystyle \frac{x^2y^2}{3}}{\displaystyle \frac{2}{\pi }}x^3y{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}{\displaystyle \frac{\mathrm{csch}({\displaystyle \frac{n\pi y}{x}})}{x^2+n^2\pi ^2}}`$ (129)
$`+`$ $`{\displaystyle \frac{2}{\pi }}xy^3{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}{\displaystyle \frac{\mathrm{csch}({\displaystyle \frac{n\pi x}{y}})}{y^2n^2\pi ^2}},`$ (130)
we obtain (with the modified minimal subtraction),
$`\mathrm{\Delta }_q`$ $`=`$ $`{\displaystyle \frac{a^2b^2}{48\pi ^2}}\mathrm{ln}{\displaystyle \frac{m_q^2}{\mu ^2}}`$ (131)
$``$ $`{\displaystyle \frac{ab}{8\pi ^3}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})\left(\mathrm{ci}({\displaystyle \frac{2n\pi m_q^2}{a}})\mathrm{cos}({\displaystyle \frac{2n\pi m_q^2}{a}})+\mathrm{si}({\displaystyle \frac{2n\pi m_q^2}{a}})\mathrm{sin}({\displaystyle \frac{2n\pi m_q^2}{a}})\right)`$ (132)
$`+`$ $`{\displaystyle \frac{ab}{16\pi ^3}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})(\mathrm{Ei}({\displaystyle \frac{2n\pi m_q^2}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi m_q^2}{b}})`$ (133)
$`+`$ $`\mathrm{Ei}({\displaystyle \frac{2n\pi m_q^2}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi m_q^2}{b}})).`$ (134)
Notice that the above effective action of the quark loop also develops an imaginary part when $`b0`$,
$`Im\mathrm{\Delta }_q={\displaystyle \frac{ab}{16\pi ^2}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi m_q^2}{b}}).`$ (135)
This is because the exponential integral Ei($`x`$) in (66) develops an imaginary part after the analytic continuation from $`x`$ to $`x`$. One can compare this with the imaginary part of the gluon loop (60). Remarkably the signature of the imaginary part of the quark loop is opposite to that of the gluon loop. This is due to the opposite statistics between the gluons and the quarks, which gives an overall minus sign for the quark loop in (64). This tells that the quarks should contribute a positive imaginary part to the effective action when $`b0`$, This means that, just like in QED, the electric flux of the quarks generates the pair creation and the color screening effect, rather than the pair annihilation and the color anti-screening effect. This allows the quarks to form the hadronic bound states.
Observe that in the pure magnetic and pure electric limits the above result reduces to
$`\mathrm{\Delta }_q=\{{\displaystyle \genfrac{}{}{0pt}{}{{\displaystyle \frac{a^2}{48\pi ^2}}\mathrm{ln}{\displaystyle \frac{m_q^2}{\mu ^2}}{\displaystyle \frac{a^2}{8\pi ^4}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n^2}}\left(\mathrm{ci}({\displaystyle \frac{2n\pi m_q^2}{a}})\mathrm{cos}({\displaystyle \frac{2n\pi m_q^2}{a}})\right)b=0,}{{\displaystyle \frac{b^2}{48\pi ^2}}\mathrm{ln}{\displaystyle \frac{m_q^2}{\mu ^2}}+{\displaystyle \frac{b^2}{16\pi ^4}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n^2}}\left(\mathrm{Ei}({\displaystyle \frac{2n\pi m_q^2}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi m_q^2}{b}})\right)a=0.}}`$ (136)
The contribution of the quark loop to the effective action is plotted in Fig.5 and Fig.6 . In view of the experimental fact that $`\mathrm{\Lambda }_{\overline{MS}}250`$ Mev and $`m_q5`$ Mev, we have assumed $`m_q=0.02`$ to obtain the figures.
Notice that in the massless limit we have
$`\mathrm{\Delta }_q|_{m_q=0}`$ $``$ $`{\displaystyle \frac{N_f}{48\pi ^2}}[a^2b^2{\displaystyle \frac{6ab}{\pi }}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}(\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})\mathrm{coth}({\displaystyle \frac{n\pi a}{b}}))]\mathrm{ln}\left({\displaystyle \frac{m_q}{\mu }}\right)^2`$ (137)
$``$ $`{\displaystyle \frac{ab}{8\pi ^3}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\left[\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})\left(\mathrm{ln}({\displaystyle \frac{2n\pi \mu ^2}{a}})+\gamma \right)\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})\left(\mathrm{ln}({\displaystyle \frac{2n\pi \mu ^2}{b}})+\gamma \right)\right]`$ (138)
$`+`$ $`i{\displaystyle \frac{ab}{16\pi ^2}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi a}{b}}),`$ (139)
so that, when $`ab0`$, the effective action from the quark loop (unlike the gluon loop) becomes divergent in the massless limit. This is because here the infra-red divergence comes from the zero modes which are physical, so that the infra-red divergence can not be removed.
One could separate the divergent part from the finite part in $`\mathrm{\Delta }_q`$. We find
$`\mathrm{\Delta }_q|_{m_q=0}=\mathrm{\Delta }_{\mathrm{}}+\mathrm{\Delta }_{\mathrm{fin}},`$ (140)
where
$`\mathrm{\Delta }_{\mathrm{}}`$ $``$ $`{\displaystyle \frac{N_f}{48\pi ^2}}[a^2b^2{\displaystyle \frac{6ab}{\pi }}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}(\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})\mathrm{coth}({\displaystyle \frac{n\pi a}{b}}))]\mathrm{ln}\left({\displaystyle \frac{m_q}{\mu }}\right)^2`$ (141)
$`+`$ $`{\displaystyle \frac{ab}{16\pi ^3}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}\left(\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})+\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})\right)\mathrm{ln}{\displaystyle \frac{a}{b}}`$ (142)
$`+`$ $`i{\displaystyle \frac{ab}{16\pi ^2}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi a}{b}}),`$ (143)
$`\mathrm{\Delta }_{\mathrm{fin}}`$ $`=`$ $`{\displaystyle \frac{ab}{16\pi ^2}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\left(\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})\right)`$ (145)
$`\left(2\gamma +\mathrm{ln}({\displaystyle \frac{2n\pi \mu ^2}{a}})+\mathrm{ln}({\displaystyle \frac{2n\pi \mu ^2}{b}})\right).`$
This tells that one has to keep $`m_q`$ finite in the evaluation of the effective action of QCD when $`ab0`$, to avoid the infra-red divergence of the quark loop.
But notice that, when $`ab=0`$, the logarithmic divergence in (69) disappears due to the following identity ,
$`{\displaystyle \frac{6ab}{\pi }}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\left(\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})\right)=a^2b^2.`$ (146)
So in this case (i.e., when $`ab=0`$) $`\mathrm{\Delta }_q`$ becomes finite even in the massless limit,
$`\mathrm{\Delta }_q|_{m_q=0}=\{{\displaystyle \genfrac{}{}{0pt}{}{{\displaystyle \frac{a^2}{48\pi ^2}}N_f(\mathrm{ln}{\displaystyle \frac{a}{\mu ^2}}c_q)b=0}{{\displaystyle \frac{b^2}{48\pi ^2}}N_f(\mathrm{ln}{\displaystyle \frac{b}{\mu ^2}}c_q)+i{\displaystyle \frac{b^2}{96\pi }}a=0,}}`$ (147)
where
$`c_q=\gamma +\mathrm{ln}(2\pi )+{\displaystyle \frac{6}{\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n^2}}\mathrm{ln}n=2.98504\mathrm{}`$ (148)
Notice that we can also obtain the above result from (68) by making the massless limit of the quarks.
## VII Effective Action and Duality
With the above analysis we can sum up the gluon and quark contributions and obtain the following final effective action of $`SU(2)`$ QCD,
$`_{eff}`$ $`=`$ $`_0+\mathrm{\Delta }_g+\mathrm{\Delta }_q`$ (149)
$`=`$ $`{\displaystyle \frac{a^2b^2}{2g^2}}+{\displaystyle \frac{3a^23b^2}{8\pi ^3}}{\displaystyle \frac{11a^2+b^2}{48\pi ^2}}\mathrm{ln}({\displaystyle \frac{2a}{\mu ^2}})+{\displaystyle \frac{11b^2+a^2}{48\pi ^2}}\mathrm{ln}({\displaystyle \frac{2b}{\mu ^2}})+{\displaystyle \frac{a^2b^2}{48\pi ^2}}\mathrm{ln}{\displaystyle \frac{m_q^2}{\mu ^2}}`$ (150)
$`+`$ $`{\displaystyle \frac{ab}{4\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\left(\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\right)(\mathrm{ci}(2n\pi )\gamma )`$ (151)
$`+`$ $`{\displaystyle \frac{ab}{8\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}[\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})(\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})\mathrm{Ei}({\displaystyle \frac{2n\pi b}{a}})+\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})\overline{\mathrm{Ei}}({\displaystyle \frac{2n\pi b}{a}}))`$ (152)
$``$ $`\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})(\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}})\mathrm{Ei}({\displaystyle \frac{2n\pi a}{b}})+\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}})\overline{\mathrm{Ei}}({\displaystyle \frac{2n\pi a}{b}}))]`$ (153)
$``$ $`{\displaystyle \frac{ab}{4\pi ^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}\left(\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{ln}({\displaystyle \frac{n\pi \mu ^2}{a}})\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{ln}({\displaystyle \frac{n\pi \mu ^2}{b}})\right)`$ (154)
$`+`$ $`{\displaystyle \frac{ab}{8\pi ^3}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi b}{a}})\left(\mathrm{ci}({\displaystyle \frac{2n\pi m_q^2}{a}})\mathrm{cos}({\displaystyle \frac{2n\pi m_q^2}{a}})+\mathrm{si}({\displaystyle \frac{2n\pi m_q^2}{a}})\mathrm{sin}({\displaystyle \frac{2n\pi m_q^2}{a}})\right)`$ (155)
$`+`$ $`{\displaystyle \frac{ab}{16\pi ^3}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})\left(\mathrm{Ei}({\displaystyle \frac{2n\pi m_q^2}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi m_q^2}{b}})+\mathrm{Ei}({\displaystyle \frac{2n\pi m_q^2}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi m_q^2}{b}})\right)`$ (156)
$``$ $`i{\displaystyle \frac{a^2+11b^2}{96\pi }}i{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n}}(\mathrm{csch}({\displaystyle \frac{n\pi b}{a}})\mathrm{exp}({\displaystyle \frac{2n\pi b}{a}})`$ (157)
$`+`$ $`\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi a}{b}}))+i{\displaystyle \frac{ab}{8\pi ^2}}{\displaystyle }_{n=1}^{\mathrm{}}{\displaystyle \frac{(1)^n}{n}}\mathrm{csch}({\displaystyle \frac{n\pi a}{b}})`$ (158)
$`+`$ $`i{\displaystyle \frac{ab}{16\pi ^2}}N_f{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n}}\mathrm{coth}({\displaystyle \frac{n\pi a}{b}})\mathrm{exp}({\displaystyle \frac{2n\pi m_q^2}{b}}).`$ (159)
From this we finally obtain Fig.7 and Fig.8, which describe the real and imaginary parts of the total effective action of $`SU(2)`$ QCD. Remember that we have assumed $`\alpha _s=1,\mu =1,m_q=0.02,`$ and $`N_f=2`$ to obtain the figures.
A truly remarkable feature of our effective action is that it is manifestly invariant under the dual transformation
$`aib,bia.`$ (160)
In fact $`\mathrm{\Delta }_g`$ and $`\mathrm{\Delta }_q`$ independently can be shown to be invariant under the dual transformation. This tells that, as a function of $`z=a+ib,`$ the effective action of QCD is invariant under the reflection from $`z`$ to $`z`$. To establish the duality in the effective action it is important to realize that the argument of the special functions ci($`x`$), si($`x`$), and Ei($`x`$) changes the signature under the dual transformation. So the dual transformation automatically involves the analytic continuation of $`x`$ to $`x`$, and one must figure out how to make the correct analytic continuation under the dual transformation. Here again the causality becomes the guiding principle. Observe that the causality requires that under the dual transformation we have
$`aiฯต`$ $`i(b+ฯต)`$ (161)
$`b+ฯต`$ $`i(aiฯต),`$ (162)
so that we must have
$`{\displaystyle \frac{a}{b}}iฯต`$ $`{\displaystyle \frac{b}{a}}iฯต`$ (163)
$`{\displaystyle \frac{b}{a}}+iฯต`$ $`{\displaystyle \frac{a}{b}}+iฯต.`$ (164)
With this it is straightforward to establish that each of $`\mathrm{\Delta }_1+\mathrm{\Delta }_2,\mathrm{\Delta }_3`$, and $`\mathrm{\Delta }_q`$ separately is invariant under the duality.
From the physical point of view the existence of the duality is not so surprising. In fact one should have expected this, because the integral expression (47) of the effective action evidently has this duality. The really remarkable fact is that this duality is borne out from our calculation of the effective action. It must be emphasized that this is a non-trivial feat, because the effective actions from the earlier calculations (including the Savvidy-Nielsen-Olesen effective action) have no such duality. This means that the duality provides a powerful tool to check the consistency of the one loop effective action. In particular one can use the duality to check the imaginary part of the effective action, because the duality intrinsically involves the analytic continuation, and thus mixes the real and imaginary parts of the effective action in a non-trivial manner. We emphasize that the consistency with the duality of our effective action provides another (a third) independent argument which supports our results.
Notice that this type of electric-magnetic duality has also been established recently in the effective action of QED . This tells that our duality is a generic feature of the gauge theories, both Abelian and non-Abelian.
It should be emphasized that it is exactly the same interaction which provided the asymptotic freedom that is responsible for the confinement. The underlying dynamics for both the asymptotic freedom and the magnetic confinement is the magnetic moment interaction. In this interaction the only difference between the quark and the gluon is the color charge, the gyromagnetic ratio, and the statistics. It has been well known that the gluon contributes positively but the quark contributes negatively to the asymptotic freedom. In this paper we have argued that this was because the gluon generates the anti-screening effect but the quark generates the screening effect. Furthermore we have proved that exactly the same physics ensures the monopole condensation and the confinement. A simple consequence of this is that we need exactly the same maximum number of the quark flavor, $`N_f=10`$ for $`SU(2)`$ and $`N_f=16`$ for $`SU(3)`$, to guarantee both the asymptotic freedom and the confinement.
Notice that with (23) the effective Lagrangian can be approximated as
$`_{eff}`$ $``$ $`{\displaystyle \frac{1}{4}}\stackrel{}{H}_{\mu \nu }^2{\displaystyle \frac{1}{2}}m^2\stackrel{}{C}_\mu ^2`$ (165)
$`=`$ $``$ $`{\displaystyle \frac{m^2}{2g^2}}(_\mu \widehat{n})^2{\displaystyle \frac{1}{4g^2}}(_\mu \widehat{n}\times _\nu \widehat{n})^2,`$ (166)
near the trivial vacuum. This of course is nothing but the Skyrme-Faddeev Lagrangian which allows the topological knot solitons as the classical solutions . This shows that there exists a deep connection between the generalized non-linear sigma model and QCD, which is very interesting.
## VIII Discussion
In this paper we have demonstrated the existence of a genuine dynamical symmetry breaking in QCD triggered by the monopole condensation. Furthermore we have established that the monopole condensation describes the stable unique vacuum of QCD. We were able to do this by calculating the one loop effective action of $`SU(2)`$ QCD. There have been earlier attempts to calculate the effective action, but these attempts have not produced a satisfactory result. We have obtained a compact expression of the effective action with an arbitrary background field. In the special cases in which the compact expressions of the effective action were available (in particular in the pure magnetic background), our result differs from the earlier results. The main difference with the earlier attempts was the controversial imaginary part in the effective action in the pure magnetic background. This has made the Savvidy-Nielsen-Olesen vacuum unstable. This assertion on the instability of the vacuum has never been seriously challenged, nor convincingly revoked. Our analysis tells that this assertion is based on the improper infra-red regularization in the evaluation of the effective action, as Schanbacher first argued . Indeed with a proper infra-red regularization we have shown that the QCD vacuum is not only stable, but is unique, made of the monopole condensation. We have provided three independent arguements to support our conclusion.
It is truly remarkable (and surprising) that the principles of the quantum field theory allow us to demonstrate the confinement within the framework of QCD. This appears against the conventional wisdom. Recently increasing number of people have been questioning the ability of the quantum field theory to provide the confinement in QCD. Indeed the failure to establish the confinement within the framework of QCD has encouraged the idea that perhaps a supersymmetric generalization of QCD may be necessary to ensure the confinement . Our analysis shows that this is not necessary after all. The QCD by itself is able to generate the confinement. What made this possible for us is the realization that we must treat the tachyonic bound states in the magnetic background and the anti-causal propagating states in the electric background which exist in the long distance region as the unphysical modes, and exclude them from the physical spectrum. In particular, we must exclude these unphysical modes from the calculation of the effective action with a proper infra-red regularization. Only this exclusion of the unphysical modes can give us a consistent theory of QCD. The fact that one could establish the dynamical symmetry breaking and the confinement in QCD within the framework of the existing quantum field theory should be interpreted as a triumph, indeed a most spectacular triumph, of the quantum field theory itself.
We conclude with the following remarks:
1) It should be emphasized that our analysis is based on the gauge independent decomposition (1) of the non-Abelian gauge potential to the restricted potential $`\widehat{A}_\mu `$ and the valence gluon $`\stackrel{}{X}_\mu `$. This is made possible with our Abelian projection . The restricted potential satisfies the full non-Abelian gauge degrees of freedom and forms a non-Abelian connection space of its own, in spite of the fact that it describes only the dual dynamics. The valence gluon forms a gauge covariant vector field, and has the gauge invariant magnetic moment interaction (10) with the restricted potential. And it is this interaction that is responsible for both the asymptotic freedom and the confinement. The existence of the gauge independent decomposition of the non-Abelian potential and a self-consistent restricted QCD has been known for more than twenty years , but its physical significance appears to have been appreciated very little so far. Now we emphasize that it is this decomposition which allows us to obtain the effective action of QCD. In particular, it is this decomposition which shows that the vacuum condensation is indeed made of the monopole condensation. Many of the earlier approaches had the critical defect that the decomposition of the non-Abelian gauge potential to the $`U(1)`$ potential and the charged vector field was not gauge independent, which has made these approaches controversial.
2) One might question (legitimately) the validity of the one loop approximation, since in the infra-red limit the non-perturbative effect is supposed to play the essential role in QCD. Our attitude on this issue is that QCD can be viewed as the perturbative extension of the topological field theory described by the restricted QCD, so that the non-perturbative effect in the low energy limit can effectively be represented by the topological structure of the restricted gauge theory. This is reasonable, because the large scale structure of the monopole topology naturally describes the long range behavior of the theory. In fact one can argue that it is the restricted potential that contributes to the Wilson loop integral, which provides a natural confinement criterion in QCD . So we believe that our monopole background automatically takes care of the essential feature of the non-perturbative effect. Of course, one could go further and try to calculate the two loop effective action , which certainly will improve our one loop correction. But this improvement is not expected to give any qualitative change, so that the generic features of the one loop effective action and the underlying physics will remain the same.
3) There have been two competing proposals for the correct mechanism of the confinement in QCD, the one emphasizing the role of the instantons and the other emphasizing that of the monopoles. Our analysis strongly favors the monopoles as the physical source for the confinement. It provides a natural dynamical symmetry breaking, and generates the mass gap necessary for the confinement in QCD. Notice that the multiple vacua, even though it is an important characteristics of the restricted gauge theory, did not play any crucial role in our calculation of the effective action. Moreover our result shows that it is the monopole condensate, not the $`\theta `$-vacuum, which describes the physical vacuum of QCD.
Although we have concentrated to $`SU(2)`$ QCD in this paper, it must be clear from our analysis that the magnetic condensation is a generic feature of the non-Abelian gauge theory. A more detailed discussion which supports our conclusions and the generalization of our result to $`SU(3)`$ will be presented in a forthcoming paper .
Acknowledgements
One of the authors (YMC) thanks S. Adler, L. Faddeev, and A. Niemi for the fruitful discussions, and Professor C. N. Yang for the continuous encouragements. The other (DGP) thanks Professor C. N. Yang for the fellowship at Asia Pacific Center for Theoretical Physics, and appreciates Haewon Lee for numerous discussions. The work is supported in part by the BK21 project of Ministry of Education.
|
warning/0006/cond-mat0006257.html
|
ar5iv
|
text
|
# New features of dislocation structures arising from lattice discreteness.
## Abstract
New aspects of a relation between lattice and dislocation structures are examined within a physically transparent theoretical scheme. Predicted features originating from the lattice discreteness include: (i) multiple core dislocation structures and (ii) their dependence on the position of the dislocation axis. These effects, which in principle can be observed directly and may also manifest themselves in dislocation motion or/and transformation (cross-slip) characteristics, are very general and present in any crystal in which they may be more or less pronounced depending on the material.
It is widely accepted that such defects as dislocations significantly influence a number of properties in real materials. Thus, understanding the relationship between lattice and dislocation structures is one of the fundamental problems of materials physics. Despite recent developments of powerful atomistic simulation techniques, up to now our understanding of the relation between lattice and dislocation structures is based on results obtained within the framework of the Peierls-Nabarro (PN) model. This model has provided both language for interpretation experimental/theoretical results and simple relations between dislocation properties and lattice discreteness characteristics (periodicity, symmetry, etc.) . This remarkable breakthrough in understanding how dislocation properties are related with lattice characteristics became possible due to two features of this model: (i) its high tractability and (ii) its combined different length scale descriptions.
The combined โcontinuum/atomisticโ descriptions in the PN model follow clearly from the structure of the energy functionals , $`E_{tot}`$, of the dislocation displacement distribution $`\stackrel{}{u}(x)`$ (here $`x`$ is a distance from the dislocation axis in the slip plane)
$`E_{tot}(\stackrel{}{u}(x))=E_{el}(\stackrel{}{u}(x))+E_{mis}(\stackrel{}{u}(x))`$ (1)
with a linear elastic ($`E_{el}`$ ) and a non-linear atomistic misfit energy term ($`E_{mis}`$)
$`E_{mis}=h{\displaystyle \underset{n}{}}\mathrm{\Phi }(\stackrel{}{u}(nhl)),`$ (2)
where $`\mathrm{\Phi }(\stackrel{}{u}(x))`$ is a periodic energy profile which is often approximated by the so-called generalized stacking fault energy (GSF) or $`\gamma `$-surface . Indeed, the $`E_{mis}`$ term represents the most apparent and important lattice properties \- discreteness/periodicity/symmetry - and allows one to investigate within the PN model the relation between lattice and dislocation properties. $`E_{mis}`$ can be expanded in a Fourier series as
$`E_{mis}=E_{mis}^0+h{\displaystyle \underset{s=1}{\overset{s=\mathrm{}}{}}}J_s\mathrm{cos}{\displaystyle \frac{2\pi sl}{h}}`$ (3)
where $`E_{mis}^0`$ $`=\underset{\mathrm{}}{\overset{\mathrm{}}{}}\mathrm{\Phi }(\stackrel{}{u}(x))๐x`$ is independent of the position of the dislocation axis $`l`$ and terms which are oscillatory with $`h`$ being a repeat distance normal to the dislocation line and $`n`$ an integer number that counts atomic rows in the same direction.
Now, a minimization of $`E_{tot}(\stackrel{}{u}(x))`$ allows one to find the equilibrium dislocation structure, $`\stackrel{}{u}(x)`$. In order to perform this minimization in analytic form, a critical approximation $`E_{mis}=E_{mis}^0`$ has been made . This โcontinuumโ approximation in representing the misfit energy - which is supposed to describe lattice discreteness - results in an obvious inconsistency which has been a subject of debate for years, mainly in the context of Peierls stress determinations . Thus, an interesting and fundamental issue arises - if consistently represented in the atomistic interaction energy, how will lattice discreteness be manifested in the structure of dislocations ?
Several recent attempts to overcome this inconsistency resulted in purely numerical procedures . Thus, one of the most advantageous features of the PN model - high tractability and transparency - has been sacrificed. These authors also focused on the Peierls stress determination and demonstrated that indeed the discrete representation of the misfit energy brings theoretical estimates much closer to experimental results.
In this Letter, using a physically transparent solution of the PN model with a consistent discrete representation for the misfit energy, we examine how lattice discreteness may influence dislocation structure. This allows us to predict new generic features of the dislocation structure that are independent of the PN model assumptions, and driven by lattice discreteness such as multiple core structures and their dependence on the position of the dislocation axis.
To determine dislocation structure, we perform a minimization of the total energy functional, Eq. 1, with a discrete representation of the misfit energy, Eq.2, using trial functions, $`\stackrel{}{u}(x)`$, defined from the Laurent expansion of their derivatives $`\rho _\beta (x)`$
$`\rho _\beta (x)={\displaystyle \frac{du_\beta (x)}{dx}}=Re{\displaystyle \underset{k=1}{\overset{N}{}}}{\displaystyle \underset{n=1}{\overset{p_k}{}}}{\displaystyle \frac{A_{nk}^\beta }{(xz_k^\beta )^n}},`$ (4)
where N is the maximal number, $`p_k`$ is the maximal order of the poles $`z_k^\beta `$ and $`A_{nk}^\beta `$ are expansion coefficients. It is important to note that, by definition, these trial functions provide a minimum of $`E_{tot}`$ for an arbitrary $`\mathrm{\Phi }(u)`$ potential (not only sinusoidal, as trial functions used in to parameterize total energy functional in the convenient form) in case of the โcontinuumโ approximation, $`E_{mis}=E_{mis}^0`$ . This choice of trial functions not only provides good accuracy and stability of the minimization procedure but also allows one to express $`E_{tot}`$ through parameters describing the dislocation structure. Indeed, the poles $`z_k^\beta =x_k^\beta +i\omega _k^\beta `$ have a clear meaning:$`x_k^\beta `$ gives the position and $`\omega _k^\beta `$ gives the width of the partials for the screw ($`\beta `$=1) and edge ($`\beta `$=2) components of the displacement in the partial cores. For example, for the ordinary dislocations dissociated into two Shockley partials, $`x_k^\beta =l\pm d^\beta /2`$, where d is the partials separation and $`l`$ gives the position of the whole ordinary dislocation center. In the general case with these trial functions, $`E_{tot}`$ can be presented as a numerical function of geometrical parameters, in particular, for an ordinary dislocation as a function of the set of parameters ($`\{๐ \}=\{d,\omega ,l\}`$) describing the dislocationโs structure ($`d,\omega `$) and its position in the lattice ($`l`$).
As examples, we consider ordinary dislocations for fcc metal, Ir, and an ordered alloy, CuAu, with L1<sub>0</sub> structure. In these materials, this type of dislocation normally splits into two Shockley partials and represents a very typical example of dislocation structures. To illustrate graphically minimization of $`E_{tot}`$, let us introduce elastic ($`F_{el}(d)`$) and misfit ($`F_{mis}(d)`$) generalized forces which are defined as, $`F_{el}(d)=E_{el}(d)/d`$ (the sign is chosen for convenience) and $`F_{mis}(d)=E_{mis}(d)/d`$ (here and further we drop the $`\beta `$ index since $`d^1d^2`$). In this definition, for a given $`d`$ other geometrical parameters from the complete set $`\{๐ \}`$ are taken to be such that they minimize $`E_{tot}`$. Obviously, in this case the intersection of $`F_{el}(d)`$ and $`F_{mis}(d)`$ gives a partial separation $`d`$ which corresponds to the minimum of Eq.1, provided that the second derivatives are positive.
The generalized forces calculated according to this definition using ab-initio $`\gamma `$-surfaces (see for details) in the case of the screw orientation of the unit dislocation for Ir and CuAu are presented in Fig.1 (a,c). For comparison, we also determine generalized forces for dislocations with simple model density distribution displacements composed of two delta functions, $`\rho (x)=b_1\delta (x+d/2)+b_2\delta (xd/2)`$ (see Fig. 1 (b, d)). In this case, we have a step function shaped dislocation for which $`E_{tot}`$ has a very simple form,
$`E_{tot}=Hln({\displaystyle \frac{1}{d}})+\gamma _{isf}d`$ (5)
and corresponding generalized forces $`F_{el}^{step}=H/d`$ and $`F_{mis}^{step}=\gamma _{isf}`$, where $`\gamma _{isf}`$ is the intrinsic stacking fault energy and H is a so-called prelogarithmic factor (see for example ). Interestingly enough, for this model type of dislocation the well-known simple relation between the equilibrium partials separation and stacking fault energy, $`d=H/\gamma _{isf}`$ , can be easily recovered from the functional dependence in Eq. 5. As can be seen in Fig.1, in the limit of large separation distances ($`d>>\omega `$), the PN model generalized forces defined within the โcontinuumโ approximation approach those for the step function shaped dislocation.
We now focus on how the oscillatory part of the misfit energy, usually neglected in the PN model analysis, affects dislocation structure. Remarkably enough, there are not only oscillations with $`l`$ which can be expected from the Fourier expansion of the misfit energy, Eq. 3, but also oscillations with partials separation $`d`$ for fixed $`l`$. As can be seen in Fig. 1, for Ir, despite the rather small amplitude of these oscillations, the effect of the misfit energy discrete representation is quite visible since the intersection of the generalized forces happens to be in the area where $`d`$ dependence of the F<sub>el</sub> is rather weak and so solutions are affected most by the F<sub>mis</sub> oscillations. Finally, for CuAu the effect is dramatic since the amplitude of the oscillations is very large. It is significant that the discrete representation of the misfit energy not only changes quantitatively parameters of the dislocation structure (as in the case of Ir) but may also result in qualitatively new effects (as in the case of CuAu). As is evident from Fig.1, one such general effect, which is independent of the PN model approximations, is the appearance of multiple stable dislocation core configurations. Indeed, conclusions we draw in this study about the possibility of multiple core configurations are based on generic features of the $`F_{el}(d)`$ and $`F_{mis}(d)`$ dependencies (see Fig.1) which follow from the general physics of the linear (elastic part) and non-linear (misfit part) lattice response .
Moreover, within the proposed scheme it is possible to derive the following convenient and physically transparent form for the energy functional, Eq.1,
$`E_{tot}=E_{tot}^0(d,\omega )+A(d,\omega )\mathrm{cos}{\displaystyle \frac{2\pi l}{h}}\mathrm{cos}{\displaystyle \frac{\pi d}{h}}`$ (6)
Here the $`l`$ independent first term $`E_{tot}^0=E_{el}+E_{mis}^0`$ is the energy in the โcontinuumโ approximation and the second term has an explicit oscillatory dependence both on $`l`$ and $`d`$.
This form reveals that the appearence of the above features associated with lattice discreteness are dependent on the relative contribution of the energies represented by the $`E_{tot}^0`$ and the oscillatory terms. In turn, the character of $`E_{tot}^0(d)`$ dependence and corresponding generalized forces is driven by the competition of the partials attraction described by $`E_{mis}^0`$ ( term which is dependent upon the $`\gamma `$-surface energetic characteristics and for large $`d`$, it can be well approximated by Eq. 5) and the elastic repulsion ($`E_{el}`$ which is dependent on the elastic constants and in the limit $`d>>\omega `$, has a simple dependence, see Fig.1). The influence of the oscillatory term is predetermined by its amplitude $`A(d,\omega )`$ which according to our analysis is strongly dependent on characteristics of the $`\gamma `$-surface .
It is important that Eq. 6 describes dislocation energetics for a wide range of $`d`$ and $`\omega `$ and correspondingly a complex interdependence of all geometrical parameters ($`d`$, $`\omega `$ and $`l`$). Features of dislocation structure originating in this interdependence of geometrical parameters and their impact on dislocation energetics can be seen from the calculated $`d`$ and $`l`$ dependencies of the $`E_{tot}`$ in Fig.2. Indeed, proof that there are can be more than one stable dislocation core configuration (CuAu) can be seen in Fig.2(a). Next, the dependence of the partials separation on the position of the dislocation axis in the lattice is clearly seen in Fig.2(b). A comparison of the $`d(l)`$ dependencies determined within the โcontinuumโ approximation and with a discrete representation for $`E_{mis}`$ makes it evident that lattice discreteness is the origin of the variation in equilibrium dislocation structure depending on the position of the dislocation axis (core โrelaxationโ). This variation may result in changes in the number of stable core configurations and abrupt transitions between them (as for CuAu, see Fig.2(a,b)).
These predicted features may have a profound impact on dislocation energetics. As can be seen in Fig.2(c), the dependence of the dislocation structure on the position of the dislocation axis may not only lower significantly the Peierls barrier (as also have been found in ) but may even modify the shape of the Peierls potential. It is remarkable that core โrelaxationโ along with the existence of the multiple core configurations (the case of CuAu) adds a new feature to the Peierls potential - an additional minimum - which according to the model analysis may result in characteristic changes of the temperature dependence of the yield stress. As can be seen clearly in Fig.2, the unusual shape of the Peierls potential in CuAu originates in the abrupt transitions between two stable core configurations, โ1โ and โ2โ.
We find that in addition to the known dislocation structure features, lattice discreteness is the origin of (i) multiple core configurations and (ii) their dependence on the position of the dislocation axis. Combination of these effects may result in rather complex variations of the dislocation structure over the crystal including changes in the number of stable core configurations and transitions between them. As a result, one may have to consider a distribution of the core configurations in a crystal under ambient conditions rather than one characteristic core structure which determines dislocation motion or cross-slip properties.
As follows from our analysis, predicted features of dislocation structure originating in lattice discreteness are always present in crystals, but as we demonstrate with Ir and CuAu as examples, they appear more or less pronounced depending on characteristics of the given material. These fundamental characteristics can be identified within the proposed theoretical analysis primarily due to its tractability and physical transparency . While these features, namely multiple core configurations, can be directly verified, in principle, in high resolution electron microscopy experiments, they may also reveal themselves indirectly in low temperature internal friction experiments and in mechanical properties which depend on elementary processes that are sensitive to the dislocation structure. Among such processes, we would emphasize cross-slip, where dislocation core structure and its changes under local stress may play an important role.
Work at Northwestern University was supported by the Air Force Office of Scientific Research (Grant No. F49620-95-1-0189) and at UCB by the Office of Basic Energy Science, Division of Materials Science, of the U.S. Department of Energy under contract No. DE-AC04-94AL85000.
|
warning/0006/cond-mat0006289.html
|
ar5iv
|
text
|
# Charged colloids at low ionic strength: macro- or microphase separation?
## Abstract
Phase separation in charged systems may involve the replacement of critical points by microphase separated states, or charge-density-wave states. A density functional theory for highly charged colloids at low ionic strength is developed to examine this possibility. It is found that the lower critical solution point is most susceptible to microphase separation. Moreover the tendency can be quantified, and related to the importance of small ion entropy in suppressing phase separation at low added salt. The theory also gives insights into the colloid structure factor in these systems.
There has been much interest recently in statistical physics in charged soft matter systems. Whilst much of this is biologically inspired (eg DNA condensation ), there has also been a long standing controversy in the colloid science community over the anomalous behaviour of charge stabilised colloidal suspensions at low ionic strengths . Recently , it has been suggested that the anomalies in these systems may be understood in terms of a miscibility gap which is the analogue of the vapour-liquid coexistence in the restricted primitive model (RPM) of 1:1 electrolytes . Arguably a theoretical consensus is emerging, although there remain a number of competing theories .
In all these examples, the crucial role of the counterions should not be underestimated. For the case of charged colloidal suspensions, for example, the theories show that the overall phase stability is almost entirely due to the entropy of the counterions . The same effect can be said to underpin the solubility of many water-soluble polymers . The basic point is that bulk phase separation in a charged system must be into electrically neutral phases. If this involves significant fractionation of counterions, an entropic penalty will be incurred which tends to suppress phase separation.
Clearly though, if bulk phase separation is suppressed, a possibility still exists to undergo *microphase separation*, where electroneutrality can be broken locally. Critical points are particularly susceptible to this, as first shown by Nabutovskii, Nemov and Peisakhovich (NNP) using a Landau-Ginzburg theory . Consider density fluctuations at wavevector $`q`$. At $`q0`$, fluctuations are restricted to elecrically neutral combinations, but for $`q>0`$ fluctuations can violate electroneutrality increasingly easily. Thus one might expect some softening of the modes. Indeed, if the only terms to $`O(q^2)`$ come from the long range Coulomb law, the analysis below implies that all partial structure factors have a *minimum* at $`q=0`$. Since the $`q=0`$ partial structure factors diverge as one approaches a critical point, this suggests there must exist regions around critical points where a divergence at $`q>0`$ occurs first, indicative that the critical behaviour is preempted by microphase separation. However, there are often other terms arising at $`O(q^2)`$ from elsewhere which destroy the phenomenon. A closely analogous microphase separation for polyelectrolytes in poor solvents has also been examined , but in the present study microphase separation is driven purely by electrostatic effects.
Let us start by constructing a simplified but physically motivated model for the anomalous behaviour in charged colloidal suspensions. Consider the macroions as spheres of charge $`Z`$, diameter $`2a`$, and number density $`\rho _\mathrm{M}`$ (volume fraction $`\varphi =4\pi a^3\rho _\mathrm{M}/3`$). There are small counterions and coions at number densities $`\rho _{}`$ and $`\rho _+`$ respectively. The solvent is a dielectric continuum. Without loss of generality, I suppose the small ions are univalent, and there is only one species of counterion . Since the coions come from added salt, it will be convenient to write $`\rho _+=\rho _\mathrm{S}`$. Overall, the system is electrically neutral and $`\rho _{}=\rho _\mathrm{S}+Z\rho _\mathrm{M}`$, but $`\delta \rho _\pm `$ will be retained for fluctuations.
Each macroion polarises the surrounding electrolyte solution, and becomes surrounded by a โdouble layerโ. It has been shown by many workers that the self energy of the macroion with its double layer, in Debye-Hรผckel theory, is $`(Z^2l_\mathrm{B}kT/2a)\times h(\kappa a)`$. In this $`l_\mathrm{B}=e^2/ฯตkT`$ is the Bjerrum length, the function $`h(x)=1/(1+x)`$ , and the Debye screening length, $`\kappa ^1`$, is given by $`\kappa ^2=8\pi l_\mathrm{B}\rho _\mathrm{I}`$ where $`2\rho _\mathrm{I}=\rho _{}+\rho _+=Z\rho _\mathrm{M}+2\rho _\mathrm{S}`$ is (twice) the ionic strength. This energy has a well known interpretation: it corresponds exactly to a *spherical capacitor*, charged $`\pm Ze`$, with one plate at the macroion surface and the second a distance $`\kappa ^1`$ away .
The simplest model free energy based on this is
$`F/VkT=\rho _\mathrm{S}\mathrm{log}\rho _\mathrm{S}+(\rho _\mathrm{S}+Z\rho _\mathrm{M})\mathrm{log}(\rho _\mathrm{S}+Z\rho _\mathrm{M})`$ (1)
$`+\rho _\mathrm{M}\mathrm{log}\rho _\mathrm{M}+\rho _\mathrm{M}(Z^2l_\mathrm{B}/2a)h(\kappa a).`$ (2)
The first three terms are the ideal terms, and the last is the self energy of macroions at number density $`\rho _\mathrm{M}`$. The most important omission from this is the contribution from the macroion-macroion interactions. Whilst this plays a significant role in structuring the macroions, it has been demonstrated elsewhere that it is *less significant* than the self energy, as regards the appearance of a miscibility gap. Moreover, by leaving this contribution out of the theory, we will see quite clearly how structure can develop in the system in the absence of pair interactions.
A typical phase diagram corresponding to the above free energy is shown in Fig. 1, for $`Z=10^3`$ and $`2a=100\mathrm{nm}`$. It comprises a simple miscibility gap, limited above and below by critical solution points as the salt concentration is varied. The gap occurs at very low ionic strengths, and only appears if the charge on the macroions is sufficiently high ($`Zl_\mathrm{B}/a13.4`$ for $`2a`$ in the range 10โ1000$`\mathrm{nm}`$).
Now, the NNP scenario could occur at either critical point. To examine this therefore, I construct a density functional theory to correspond to the free energy introduced above. The ideal terms become $`kTd^3๐ซ\rho _i(๐ซ)\mathrm{log}\rho _i(๐ซ)`$ ($`i=+,,\mathrm{M}`$), and I introduce the *ansatz* that the self energy generalises in the obvious way to $`d^3๐ซ\rho _\mathrm{M}(๐ซ)f_N^{\mathrm{self}}(๐ซ)`$ where $`f_N^{\mathrm{self}}=(Z^2l_\mathrm{B}kT/2a)h(\kappa a)`$ is the self energy per particle evaluated using the local ionic strength at the particle centre, $`\kappa ^2=8\pi l_\mathrm{B}\rho _\mathrm{I}(๐ซ)`$. Finally, an electrostatic contrbution has to be added: $`l_\mathrm{B}kTd^3๐ซd^3๐ซ^{}\rho _\mathrm{Z}(๐ซ)\rho _\mathrm{Z}(๐ซ^{})/|๐ซ๐ซ^{}|`$, where $`\rho _\mathrm{Z}(๐ซ)=Z\rho _\mathrm{M}(๐ซ)+\rho _+(๐ซ)\rho _{}(๐ซ)`$ is the local charge density.
To examine the stability of the system against microphase separation, expand the above density functional about the homogeneous state to quadratic order. For fluctuations at a wavevector $`q`$ this results in
$`{\displaystyle \frac{\delta F}{VkT}}={\displaystyle \frac{|\delta \rho _+|^2}{2\rho _\mathrm{S}}}+{\displaystyle \frac{|\delta \rho _{}|^2}{2(\rho _\mathrm{S}+Z\rho _\mathrm{M})}}+{\displaystyle \frac{|\delta \rho _\mathrm{M}|^2}{2\rho _\mathrm{M}}}`$ (3)
$`+{\displaystyle \frac{2\pi l_\mathrm{B}}{q^2}}|Z\delta \rho _\mathrm{M}+\delta \rho _+\delta \rho _{}|^2`$ (4)
$`+{\displaystyle \frac{Z^2l_\mathrm{B}}{2a}}[8\pi ^2l_\mathrm{B}^2a^4\rho _\mathrm{M}h_1(\kappa a)|\delta \rho _\mathrm{I}|^2`$ (5)
$`2\pi l_\mathrm{B}a^2h_2(\kappa a)(\delta \rho _\mathrm{M}\delta \rho _\mathrm{I}^{}+\delta \rho _\mathrm{M}^{}\delta \rho _\mathrm{I})]`$ (6)
The functions are $`h_1(x)=(1+3x)/(x^3(1+x)^3)`$ and $`h_2(x)=1/(x(1+x)^2)`$. From this the macroion structure factor $`S(q)=|\delta \rho _\mathrm{M}(q)|^2`$ is extracted in the standard way . The behaviour of $`S(q)`$ is examined as a function of $`\varphi `$ and $`\rho _\mathrm{S}`$, looking for the unstable regions in the $`(\varphi ,\rho _\mathrm{S})`$-plane where $`1/S(q)<0`$.
At $`q=0`$ the spinodal instability region corresponding to the free energy in Eq. (2) is recovered. For $`q>0`$ the region of instability *always expands*. This is because the $`q`$-dependence arises solely from the long range electrostatic term, thus, as alluded to above, $`S(q)`$ always has a minimum at $`q=0`$. But a clear difficulty emerges when the behaviour for large $`q`$ is examined, since the instability region expands to fill the entire plane; there is no effective penalty against microphase separation at vanishingly small wavelengths.
In fact the *ansatz* for the self energy term has omitted an obvious but crucial effect, namely one would not expect a macroion to be sensitive to variations in the local ionic strength over distances much smaller than its size. Motivated by weighted local density theories for liquids , I therefore introduce an additional *smoothing* *ansatz*. It turns out that the precise form does not matter greatly, for instance one can replace the local ionic strength by $`\overline{\rho }_\mathrm{I}(๐ซ)=d^3๐ซ^{}w(|๐ซ๐ซ^{}|)\rho _\mathrm{I}(๐ซ^{})`$ where $`w(r)`$ is a smoothing kernel of range $`a`$ , but equally one might smooth $`\kappa `$ or $`f_N^{\mathrm{self}}`$. All forms result in the appearance of extra multiplicative factors, $`w_1(qa)`$ and $`w_2(qa)`$, in the last two terms of Eq. (5). The $`w_i(qa)`$ are related to the Fourier transform of $`w(r)`$, and satisfy $`w_i1`$ at $`q0`$, $`w_i0`$ at $`q\mathrm{}`$. (More generally I expect $`h_i(\kappa a,qa)`$ such that $`h_i0`$ for $`qa\mathrm{}`$.) With this *ansatz*, progress can be made without developing a detailed theory by investigating various possibilities for $`w_i`$. The results reported below have been carried out assuming $`w_1=w_2=\mathrm{exp}(\alpha q^2a^2)`$ with $`\alpha `$ a numerical prefactor of order unity. Very similar results are obtained for $`w_i=1/(1+\alpha q^2a^2)`$.
Typical results from this modified theory are shown in Figs. 2 and 3, for $`\alpha =1`$. In Fig. 2, the spinodal instability at $`q=0`$ is recovered as before. For $`q>0`$, the instability region is again expanded in the vicinity of the lower critical point, but is now *reduced* in the vicinity of the upper critical point. For $`qa1`$ the instability disappears completely, since the self energy which drives the instability is now insensitive to short wavelength fluctuations. Below the lower critical point, therefore, there is a region (delimited by the heavy dashed line in Fig. 2) where $`S(q)`$ diverges at some $`q^{}>0`$, corresponding to the appearance of microphases.
These results are reflected in the macroion structure factors, two examples of which are shown in Fig. 3. Near the upper critical point the structure factor turns up to a maximum as $`q0`$, developing a divergence at $`q=0`$ as the (mean field) critical point is approached. Near the lower critical point a peak appears in $`S(q)`$ at $`q^{}>0`$. The peak diverges as one approaches the boundary of the microphase instability region. Note that the appearance of a peak in the macroion structure factor $`S(q)`$ is unusual because there are no direct macroion interactions in the theory as constituted above. This shows how structure can arise in a charged system independent of the existence of (effective) pair interactions.
To relate the results to possible experiments, one should of course investigate the significant contribution to the structure factor from the omitted macroion interactions. Elsewhere it is argued that, at these low ionic strengths, the macroions are effectively a one-component plasma (OCP) in the strong coupling limit , whose structure factor resembles that of hard sphere (HS) fluid close to the freezing transition . A rescaled $`S(q)`$ for HS at freezing is compared with the present calculations in the inset in Fig. 3. The additional structure arising from the self energy theory above appears to lie well within the first peak, and moreover the amplitudes match quite closely. Thus the appearance of a maximum at $`q=0`$ or a peak at very low $`q`$ may well be experimentally observable. This may be the explanation of the anomalously large $`S(q0)`$ reported recently for colloidal suspensions at low ionic strengths .
Let us turn to the effect of $`\alpha `$. Recall that $`\alpha `$ is a measure of the degree of smoothing: the range of the smoothing kernel is $`a\sqrt{\alpha }`$. All the results discussed above were at $`\alpha =1`$. If $`\alpha 0.338`$, microphases appear at the *upper* critical point too. On the other hand, if $`\alpha 1.961`$ microphase separation at the lower critical point disappears. One would expect increasing $`\alpha `$ to suppress microphase separation, since greater smoothing is bound to reduce $`q^{}`$, but the difference between the two critical points is suggestive. These critical values of $`\alpha `$ are more general than the assumed form of $`w_i(q)`$, since they only depend on the $`q^2`$ coefficient in the expansion $`w_i(q)=1\alpha q^2a^2+O(q^4)`$ . They are a measure of the susceptibility of the critical point to replacement by microphases (if $`\alpha `$ could be treated as a control variable, the critical values would correspond to Lifshitz points in the phase diagram).
This calculation sheds light on the reason for a *closed loop* miscibility gap. As $`\rho _\mathrm{S}`$ is increased, an upper critical point is expected since the self energy ceases to be strongly state point dependent for $`\kappa a1`$ or $`\rho _\mathrm{S}Z\rho _\mathrm{M}`$, where it simply shifts the macroion chemical potential (see lines in Fig. 1). As $`\rho _\mathrm{S}0`$ though, it is perhaps unexpected to encounter a *second* critical point. Its appearance appears to be connected to the small ion entropy effect discussed in the introduction. The evidence for this is twofold. Firstly, as already commented upon, the lower critical point is more susceptible to microphase separation. This is in accord with the idea that phase separation is suppressed by small ion entropy, since in microphase separation, the system gains entropy by distributing the small ions more uniformly than would be allowed if strict electroneutrality had to be satisfied at each point.
The second piece of evidence concerns the rate at which the *Donnan potential difference* $`\mathrm{\Delta }\overline{\psi }`$ vanishes as one approaches the critical point. Recall that $`\mathrm{\Delta }\overline{\psi }`$ arises because the interface can acquire a dipole moment density (per unit area). In the present system, a dipole moment density appears to arise because the jump in small ion densities is spread out more broadly than the jump in macroion densities (although this remains to be confirmed with a detailed calculation ). Remarkably, one can calculate $`\mathrm{\Delta }\overline{\psi }`$ without detailed knowledge of the structure of the interface . I find that $`\mathrm{\Delta }\overline{\psi }`$ vanishes as $`\mathrm{\Delta }\varphi /\varphi _{\mathrm{crit}}`$ as the critical points are approached, with a constant of proportionality $`8.00`$ for the upper critical solution point, and $`16.9`$ for the lower one. This again indicates the growing importance of small ion entropy (broadening the jump in small ion densities) as the lower critical point is approached.
Note that $`\mathrm{\Delta }\overline{\psi }`$ is an order parameter which strictly vanishes in symmetric models such as the RPM. Apart from general remarks by Nabutovskii and Nemov , the critical behaviour of asymmetric primitive models seems to have received much less attention than the RPM , and there may be interesting effects connected to a non-vanishing $`\mathrm{\Delta }\overline{\psi }`$.
I thank C. Holmes and M. E. Cates for useful discussions, and P. Schurtenberger for sending data prior to publication.
|
warning/0006/gr-qc0006080.html
|
ar5iv
|
text
|
# ON THE INFLUENCE OF GRAVITATIONAL RADIATION ON A GYROSCOPE
## 1 Introduction
The theoretical description and the experimental observation of gravitational radiation are among the most relevant challenges confronting general relativity.
A great deal of work has been done so far in order to provide a consistent framework for the study of such phenomenon. Also,since Weberโs pioneering work important collaboration efforts have been carried on, and are now under consideration, to put in evidence gravitational waves (see , , ,, and references therein).
It is the purpose of this work to evaluate the influence of gravitational radiation on a gyroscope.This idea is not new, in fact some years ago, Chaboyer and Henriksen put forward the possibility of detecting gravitational radiation by means of an orbital laser gyroscope. In such experiment the presence of gravitational radiation is brought out by the differential effect that radiation has on the paths of the photons in the rotating frame of reference.
In this work we shall calculate the rate of precession of a gyroscope in the field of gravitational radiation. To do so we shall use the Bondiโs formalism which has, among other things, the virtue of providing a clear and precise criterion for the existence of gravitational radiation (see also ). Namely, if the news function is zero over a time interval, then there is no radiation during that interval. Also, the present approach has the advantage of providing a very simple expression linking anโobservableโ (at least in principle) quantity (the rate of precession of a gyroscope) with the emission rate of gravitational radiation.
The formalism has as its main drawback the fact that it is based on a series expansion which could not give closed solutions and which raises unanswered questions about convergence and appropriateness of the expansion.
However since we shall assume the gyroscope to be very far from the source, we shall use in our calculations only the leading terms in the expansion of metric functions. Furthermore, since the source is assumed to radiate during a finite interval, then no problem of convergence appears .
We shall see that the leading term of the rate of precession of the gyroscope ($`\mathrm{\Omega }`$) is expressed through the news function in such a way that it will vanish if and only if there is no news (no radiation).
In the special case of the quadrupole radiation (in the linear approximation), the rate of precession may be expressed through the third time derivative of the quadrupole moment, or alternatively, through the rate of loss of the mass function.
Next we present the order ($`{\displaystyle \frac{1}{r^2}}`$) of $`\mathrm{\Omega }`$. As we shall see, it contains terms with news, together with a time dependent term not involving news. This last term represents the class of non-radiative motions discussed by Bondi and may be thought to correspond to the tail of the wave, appearing after the radiation process . The obtained expression allows for โmeasuringโ (in a gedanken experiment, at least) the wave-tail field. This in turn implies that observing the gyroscope, for a period of time from an initial static situation until after the vanishing of the news, should allow for an unambiguous identification of a gravitational radiation process.
In the next section we briefly present the Bondiโs formalism. The expression for the rate of precession of the gyroscope in the Bondi metric is calculated in sections 3 and 4, and estimates for the rate of precession in different scenarios are presented in section 5. Finally the results are discussed in the last section.
## 2 The Bondiโs Formalism
The general form of an axially symmetric asymptotically flat metric given by Bondi is
$`ds^2`$ $`=`$ $`\left({\displaystyle \frac{V}{r}}e^{2\beta }U^2r^2e^{2\gamma }\right)du^2+2e^{2\beta }dudr`$ (1)
$`+`$ $`2Ur^2e^{2\gamma }dud\theta r^2\left(e^{2\gamma }d\theta ^2+e^{2\gamma }\mathrm{sin}^2\theta d\varphi ^2\right)`$
where $`V,\beta ,U`$ and $`\gamma `$ are functions of $`u,r`$ and $`\theta `$.
We number the coordinates $`x^{0,1,2,3}=u,r,\theta ,\varphi `$ respectively. $`u`$ is a timelike coordinate such that $`u=constant`$ defines a null surface. In flat spacetime this surface coincides with the null light cone open to the future. $`r`$ is a null coordinate ($`g_{rr}=0`$) and $`\theta `$ and $`\varphi `$ are two angle coordinates (see for details).
Regularity conditions in the neighborhood of the polar axis ($`\mathrm{sin}\theta =0`$), implies that as $`\mathrm{sin}\theta >0`$
$$V,\beta ,U/\mathrm{sin}\theta ,\gamma /\mathrm{sin}^2\theta $$
(2)
each equals a function of $`\mathrm{cos}\theta `$ regular on the polar axis.
The four metric functions are assumed to be expanded in series of $`1/r`$, then using field equations Bondi gets
$$\gamma =c(u,\theta )r^1+\left(C(u,\theta )\frac{1}{6}c^3\right)r^3+\mathrm{}$$
(3)
$$U=\left(c_\theta +2c\mathrm{cot}\theta \right)r^2+\mathrm{}$$
(4)
$`V`$ $`=`$ $`r2M(u,\theta )`$ (5)
$``$ $`\left(N_\theta +N\mathrm{cot}\theta c_\theta ^24cc_\theta \mathrm{cot}\theta {\displaystyle \frac{1}{2}}c^2(1+8\mathrm{cot}^2\theta )\right)r^1+\mathrm{}`$
$$\beta =\frac{1}{4}c^2r^2+\mathrm{}$$
(6)
where letters as subscripts denote derivatives, and
$$4C_u=2c^2c_u+2cM+N\mathrm{cot}\theta N_\theta $$
(7)
The three functions $`c,M`$ and $`N`$ are further related by the supplementary conditions
$$M_u=c_u^2+\frac{1}{2}\left(c_{\theta \theta }+3c_\theta \mathrm{cot}\theta 2c\right)_u$$
(8)
$$3N_u=M_\theta +3cc_{u\theta }+4cc_u\mathrm{cot}\theta +c_uc_\theta $$
(9)
In the static case $`M`$ equals the mass of the system whereas $`N`$ and $`C`$ are closely related to the dipole and quadrupole moment respectively.
Next, Bondi defines the mass $`m(u)`$ of the system as
$$m(u)=\frac{1}{2}_0^\pi M(u,\theta )\mathrm{sin}\theta d\theta $$
(10)
which by virtue of (8) and (2) yields
$$m_u=\frac{1}{2}_0^\pi c_u^2\mathrm{sin}\theta d\theta $$
(11)
Let us now recall the main conclusions emerging from the Bondiโs approach.
1. If $`\gamma ,M`$ and $`N`$ are known for some $`u=a`$(constant) and $`c_u`$ (the news function) is known for all $`u`$ in the interval $`aub`$, then the system is fully determined in that interval. In other words, whatever happens at the source, leading to changes in the field, it can only do so by affecting $`c_u`$ and viceversa. At the light of this comment the relationship between news function and the occurrence of radiation becomes clear.
2. As it follows from (11), the mass of a system is constant if and only if there are no news.
In the next section we calculate the rate of precession of a gyroscope at rest in the frame of (1)
## 3 The gyroscopic precession
Let us start by defining the vorticity vector, which as usual is given by (in relativistic units)
$`\omega ^\alpha `$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{g}}}ฯต^{\alpha \beta \gamma \delta }u_\beta \omega _{\gamma \delta }`$ (12)
$`=`$ $`{\displaystyle \frac{1}{2\sqrt{g}}}ฯต^{\alpha \beta \gamma \delta }u_\beta u_{\gamma ,\delta }`$
where the vorticity tensor is given by
$$\omega _{\alpha \beta }=u_{[\alpha ;\beta ]}\dot{u}_{[\alpha }u_{\beta ]}$$
(13)
and $`u_\beta `$ denotes the four-velocity vector.
Now, for an observer at rest in the frame of (1), the four-velocity vector has components
$$u_\alpha =(A,\frac{e^{2\beta }}{A},\frac{Ur^2e^{2\gamma }}{A},0)$$
(14)
with
$$A\left(\frac{V}{r}e^{2\beta }U^2r^2e^{2\gamma }\right)^{1/2}$$
(15)
using (14) and
$$\sqrt{g}=e^{2\beta }r^2\mathrm{sin}\theta $$
(16)
in (12), we easily obtain
$$\omega ^\alpha =(0,0,0,\omega ^3)$$
(17)
with
$`\omega ^3`$ $`=`$ $`{\displaystyle \frac{e^{2\beta }}{2r^2\mathrm{sin}\theta }}\{2\beta _\theta e^{2\beta }{\displaystyle \frac{2e^{2\beta }A_\theta }{A}}\left(Ur^2e^{2\gamma }\right)_r`$ (18)
$`+`$ $`{\displaystyle \frac{2Ur^2e^{2\gamma }}{A}}A_r+{\displaystyle \frac{e^{2\beta }\left(Ur^2e^{2\gamma }\right)_u}{A^2}}{\displaystyle \frac{Ur^2e^{2\gamma }}{A^2}}2\beta _ue^{2\beta }\}`$
and for the absolute value of $`\omega ^\alpha `$ we get
$`\mathrm{\Omega }`$ $``$ $`(\omega _\alpha \omega ^\alpha )^{1/2}={\displaystyle \frac{e^{2\beta \gamma }}{2r}}\{2\beta _\theta e^{2\beta }2e^{2\beta }{\displaystyle \frac{A_\theta }{A}}\left(Ur^2e^{2\gamma }\right)_r`$ (19)
$`+`$ $`2Ur^2e^{2\gamma }{\displaystyle \frac{A_r}{A}}+{\displaystyle \frac{e^{2\beta }}{A^2}}\left(Ur^2e^{2\gamma }\right)_u2\beta _u{\displaystyle \frac{e^{2\beta }}{A^2}}Ur^2e^{2\gamma }\}`$
Feeding back (3)โ(6) into (19) and keeping only the leading term, we obtain
$$\mathrm{\Omega }=\frac{1}{2r}\left(c_{u\theta }+2c_u\mathrm{cot}\theta \right)+O(r^n);n>1$$
(20)
Now, since $`\mathrm{\Omega }`$ measures the rate of rotation with respect to proper time of world lines of points at rest in the frame of (1), relative to the local compass of inertia, then $`\mathrm{\Omega }`$ describes the rotation of the compass of inertia (โ the gyroscopeโ) with respect to reference particles at rest in the frame of (1) (see for detailed discussion on this point).
Therefore, up to order $`1/r`$, the gyroscope will precess as long as the system radiates ($`c_u0`$). Observe that if
$$c_{u\theta }+2c_u\mathrm{cot}\theta =0$$
(21)
then
$$c_u=\frac{F(u)}{\mathrm{sin}^2\theta }$$
(22)
which implies
$$F(u)=0c_u=0$$
(23)
in order to insure regularity conditions, mentioned above, in the neighbourhood of the polar axis ($`\mathrm{sin}\theta =0`$) . Thus the leading term in (20) will vanish if and only if $`c_u=0`$.
If the system radiates during an interval of time $`\mathrm{\Delta }u`$, then the change of orientation of the gyroscope, for that period, is given by
$$\mathrm{\Delta }\varphi =\mathrm{\Omega }\mathrm{\Delta }u\left(\frac{V}{r}e^{2\beta }U^2r^2e^{2\gamma }\right)^{1/2}$$
(24)
or, up to terms of order $`1/r`$
$$\mathrm{\Delta }\varphi \frac{1}{2r}\left(c_{u\theta }+2c_u\mathrm{cot}\theta \right)\mathrm{\Delta }u$$
(25)
Let us now consider the particular case of a quadrupole radiation in the linear approximation. If the quadrupole moment of the source is $`Q(u)`$, then it can be shown (see eqs. (86)โ(91) in ) that in the linear approximation
$$c=\frac{1}{2}Q_{uu}\mathrm{sin}^2\theta $$
(26)
and
$$m_u=\frac{2}{15}Q_{uuu}^2$$
(27)
Thus
$$\mathrm{\Omega }=\frac{1}{2r}\mathrm{sin}2\theta Q_{uuu}+O(r^n)$$
(28)
or
$$\mathrm{\Omega }=\sqrt{\frac{15}{8}}\frac{\mathrm{sin}2\theta }{r}\left(m_u\right)^{1/2}+O(r^n)$$
(29)
linking directly the rate of precession to the rate of loss of mass.
## 4 The gyroscopic preccession of order $`{\displaystyle \frac{1}{r^2}}`$
Let us now consider the next order ($`{\displaystyle \frac{1}{r^2}}`$). We easily obtain:
$`\mathrm{\Omega }`$ $`=`$ $`{\displaystyle \frac{1}{2r}}(c_{u\theta }+2c_u\mathrm{cot}\theta )`$ (30)
$`+{\displaystyle \frac{1}{r^2}}\left[M_\theta M(c_{u\theta }+2c_u\mathrm{cot}\theta )cc_{u\theta }+6cc_u\mathrm{cot}\theta +2c_uc_\theta \right]`$
Observe that the order $`{\displaystyle \frac{1}{r^2}}`$ contains, beside the terms involving $`c_u`$, a term not involving news ($`M_\theta `$). Let us now assume that initially (before some $`u=u_0=`$constant) the system is static, in which case
$$c_u=0$$
(31)
which implies , because of (9)
$$M_\theta =0$$
(32)
and $`\mathrm{\Omega }=0`$ (actually, in this case $`\mathrm{\Omega }=0`$ at any order) as expected for a static field ( for the electrovacuum case however, this may change ). Then let us suppose that at $`u=u_0`$ the system starts to radiate ($`c_u0`$) until $`u=u_f`$, when the news vanish again. For $`u>u_f`$ the system is not radiating although (in general) $`M_\theta 0`$ implying (see for example (9)) time dependence of metric functions (non-radiative motions ).
In the interval $`u`$ ($`u_0`$,$`u_f`$) the leading term of the rate of precession of the gyroscope is given by (20).
For $`u>u_f`$ there is a precession term of order $`{\displaystyle \frac{1}{r^2}}`$ describing the effect of the tail of the wave on the gyroscope. This in turn provides โobservationalโ evidence for the violation of the Huygensโs principle, a problem largely discussed in the literature (see for example ,, and references therein).
Putting aside the actual technical difficulties in performing such an experiment, it should be clear that the monitoring of the gyroscope in the interval ($`u<u_0`$, $`u>u_f`$) should, in principle, bring out, in a clear-cut way, the presence of gravitational radiation.
Finally, let us consider the particular case of a quadrupole radiation in the linear approximation. We obtain for this case (note a misprint in Eq.(87) in ).
$$\mathrm{\Omega }=\sqrt{\frac{15}{8}}\frac{\mathrm{sin}2\theta }{r}(m_u)^{1/2}+\frac{1}{r^2}(3Q_{uu}\mathrm{sin}2\theta )$$
(33)
Therefore, for $`u>u_f`$, the rate of precession is controled by the second time derivative of the quadrupole moment ($`Q_{uu}`$).
## 5 Scenarios of radiation and estimations
Let us now present some rough estimates for $`\mathrm{\Omega }`$ or $`\mathrm{\Delta }\varphi `$, in different scenarios.Before doing that, some remarks are in order.
Since our intention here is just to provide orders of magnitude of the expected effect, we shall restrain ourselves to the quadrupole radiation case. Also, although Bondi approach implies axial symmetry, it is not clear that such symmetry is present in all examples below. This may be particularly true for the case of collisions and bremsstrahlung. In the other cases, specially in those of gravitational collapse and supernovae, axial symmetry is not a too stringent condition. In the same line of arguments, it should be mentioned that in some examples, particularly in binary systems, the signal may last for a too long duration. Since the convergence of the series requires (see for details)
$$u<2r$$
(34)
then, in those examples, the source should be very far from the gyroscope, in order to assure the convergence of the series expansion. At any rate we should insist on the point that the purpose of this section is not to propose specific scenarios for eventual experiments, but just to provide , however vague, orders of magnitude of the mentioned effect.
Now, in the case of quadrupole radiation (in the linear approximation) the rate of precession can be related to the rate of loss of mass of the source through (29).Then for the change of orientation of the gyroscope $`\mathrm{\Delta }\varphi `$, during an interval of time $`\mathrm{\Delta }u`$ we obtain using (24),
$$\mathrm{\Delta }\varphi =\sqrt{\frac{15}{8}}\frac{\mathrm{sin}2\theta }{r}\left(m_u\right)^{1/2}\mathrm{\Delta }u$$
(35)
This last equation will be used to obtain estimations of $`\mathrm{\Delta }\varphi `$, whenever the rate of loss of mass (or the total radiated mass) and the time interval of radiation are available.In some examples, when there is not a characteristic time scale of the emission we give an estimate of $`\mathrm{\Omega }`$, from (29).For simplicity the numerical factor and the trigonometric term, in (35) and (29), are put equal to one.
In the last three decades, a great deal of work has been done in identifying possible sources of gravitational radiation (see , and references therein). Here we shall present a selection of some of them, keeping in mind all reserves mentioned above:
1. Bynary systems.
* For the binary system of two neutron stars proposed by Clark , at 10 Kpc, we obtain $`\mathrm{\Delta }\varphi 2.2\times 10^{15}`$. This quantity is obtained for an emission time of $`1`$ second, of a binay system emitting mass at a rate $`m_u10^{47}Joules/s`$.
* For the pulsar 1937+214 (see ), one obtains $`\mathrm{\Omega }1.23\times 10^{23}`$rad.s<sup>-1</sup>. The estimated value used for its angular velocity $`4033.8rad/s`$ leads to a power of emission around $`2\times 10^{29}watts`$, taking into account the Landauโs formula and considering a distance of $`2.5Kpc`$ for the source. Identifying the power emitted by the system with the loss of mass $`m_u`$, we obtain the given above value of $`\mathrm{\Omega }`$ .
* Similar estimations may be done for the pulsar 1913+16 (see ), knowing that the power of this pulsar is $`6.4\times 10^{23}watt`$. The estimates yield, for a distance of $`5Kpc`$, $`\mathrm{\Omega }1.13\times 10^{26}`$rad.s<sup>-1</sup>.
2. Gravitational collapse and supernovae
* According to Shapiro , during the first bounce and rebound, the efficiency of gravitational radiation never exceeds $`10^3M_{}`$,then for an event at 10Kpc with a duration of the order of 1ms (free fall time), the maximum obtained $`\mathrm{\Delta }\varphi `$ is of the order of $`3\times 10^{18}`$
* The model of collapse proposed by Wilson , assumes a radiating energy (in the form of gravitational radiation) of the order of $`10^2M_{}`$, during an interval of time of $`100M_{}`$ (in relativistic units). Then for one solar mass , at a distance of 10Kpc, one obtains $`\mathrm{\Delta }\varphi 6.4\times 10^{18}`$.
* In the strongest massive star collapse proposed by Ostriker , the emission is of the order of $`10^1`$ solar masses during $`10^1`$s, at 1 Kpc.The resulting $`\mathrm{\Delta }\varphi `$ is of the order of $`2.9\times 10^{15}`$.
* For a model of supernovae proposed by Braginsky and Rudenko , (an emission of $`10^{48}`$J/s, during 1 ms at 15Mpc) one obtains $`\mathrm{\Delta }\varphi 4.2\times 10^{21}`$.
* The characteristic parameters of a stellar collapse model proposed by Rees et al , are : an emission of $`10^{50}`$J/s during $`5\times 10^4`$s. In such a collapse there is a continuous frequency espectrum up to $`{\displaystyle \frac{1}{\tau _\rho }}`$, being $`\tau _\rho \sqrt{\pi G}\rho `$ the characteristic time of the collapse for a final state density $`\rho `$. The variation of energy $`\mathrm{\Delta }E`$ goes like $`\mathrm{\Delta }E{\displaystyle \frac{1}{30\pi }}{\displaystyle \frac{Q^2}{\tau _\rho ^5}}`$ where $`Q`$ is the quadrupole moment along the axis. For an event at 5 Kpc involving a star of $`6M_{}`$, the resulting $`\mathrm{\Delta }\varphi `$ is of the order of $`7\times 10^{17}`$.
* Finally, let us evaluate the emission of gravitational radiation in a supernovae event, leading to a neutron star, by estimating the total change of the gravitational quadrupole moment of the source. Recent estimations of the quadrupole moment of neutron stars , point to values of Q of the order of $`1.03\times 10^{37}`$kg.$`m^2`$. If we assume that the presupernovae massive star has a quadrupole moment of the order of the sun ( $`3.85\times 10^{42}`$kg.$`m^2`$) , then the total change of quadrupole moment is $`\mathrm{\Delta }Q10^{42}Kgm^2`$. Taking into account (27),(35), then the obtained $`\mathrm{\Delta }\varphi `$, for an event at distance of 1 Mpc, during $`10^7years`$ (Kelvin-Helmholtz time scale for a star like the sun), is of the order of $`2.8\times 10^{28}`$.
3. Collisions and Bremsstrahlung.
* The emission, during collision of two neutron stars, given by Wilson is of the order of $`10^3M_{}`$ during an interval time of the order of $`50M_{}`$.For an event at 10Kpc we obtain $`\mathrm{\Delta }\varphi 1.14\times 10^{18}`$.
* In the case of two black holes collision, Detweiler proposes an emission of $`10^3M_{}`$, with a collision time of the order of $`10`$M. For an event at 10Kpc this leads to $`\mathrm{\Delta }\varphi 6.4\times 10^{25}`$.
* For gravitational bremsstrahlung within the galaxy, Ostriker suggest an emission of $`10^2`$ solar masses during one second, this yields a $`\mathrm{\Delta }\varphi `$ of the order of $`3\times 10^{16}`$.
4. Other sources.
* Following a speculative discussion about the evolution of galactic nucleus, Ostriker considers the possibility of destruction of $`10^8`$ neutron stars at the center, in aproximately $`10^{6.5}`$ years, emitting $`10^2`$ solar masses in the form of gravitational radiation, an event of this kind, at 1Mpc of distance would produce a total precession of the order of $`3\times 10^{11}`$.
## 6 Conclusions
We have seen so far that a gyroscope at rest in a Bondi frame will precess (up to order $`1/r`$) as long as the system radiates, the rate of precession being given by (20).
Once the radiation stops (vanishing news) the gyroscope will continue to precess with a rate of rotation given by the second term of (30) with $`c_u=0`$ .
As can be seen from estimations done above, excluding the last example, and the example based on the evaluation of the change in the quadrupole moment, the most realistic scenarios point to $`\mathrm{\Delta }\varphi `$ โs ranging between $`10^{15}`$ and $`10^{19}`$. We ignore how far are we , with the present technology, to the accuracy required for that kind of measurement. Nevertheless, we want to stress that it has been our main purpose here, just to bring out such effects in the context of Bondi formalism.
We would like to conclude with the following comment: observe that all along our discussion we have not made reference to specific bandwiths. This is so because all quantities in the Bondi approach, are not defined with respect to any specific frecuency. In this sense eq.(11) (and all related quantities), have to be considered as integrated over all frecuencies.
## Acknowledgment
It is a pleasure to thank Prof. Bondi for interesting comments.
|
warning/0006/math0006120.html
|
ar5iv
|
text
|
# Oblique projections and Schur complements
## 1 Introduction
If $``$ is a Hilbert space with scalar product $`,`$ and $`L()`$ is the algebra of all bounded linear operators on $``$, consider the subset $`๐ฌ`$ of $`L()`$ consisting of all projections onto (closed) subspaces of $``$ and the subset $`๐ซ`$ of $`๐ฌ`$ of all orthogonal (i.e., selfadjoint) projections. Every $`Q๐ฌ๐ซ`$ is called an oblique projection. The structure of $`๐ฌ`$ and $`๐ซ`$ has been widely studied since the begining of the spectral theory. In recent times, applications of oblique projections to complex geometry , statistics , and wavelet theory , , , have renewed the interest on the subject. The reader is also referred to , .
In , there is an analytic study of the map which assigns to any positive invertible operator $`AL()`$ and any subspace $`๐ฎ`$ of $``$ the unique projection onto $`๐ฎ`$ which is selfadjoint for the scalar product $`,_A`$ on $``$ defined by $`\xi ,\eta _A=A\xi ,\eta `$ $`(\xi ,\eta )`$. In this paper we study the existence of projections onto $`๐ฎ`$ which are selfadjoint for $`,_A`$ if $`A`$ is not necesarily invertible. More precisely, if $`๐ฎ`$ is a closed subspace of $``$ and $`B:\times \text{C}\text{ }\text{ }`$ is a Hermitian sesquilinear form, consider the subsets of $`๐ฌ`$,
$$๐ฌ_๐ฎ=\{Q๐ฌ:Q()=๐ฎ\}\text{ (projections with range S) }$$
and
$$๐ฌ^B=\{Q๐ฌ:B(\xi ,Q\eta )=B(Q\xi ,\eta ),\text{ for all }\xi ,\eta \}\text{ (B-symmetric projections).}$$
The main theme of the paper is the characterization of the intersection of $`๐ฌ_๐ฎ`$ and $`๐ฌ^B`$. We shall limit our study to the case in which $`B`$ is bounded, so that, by Rieszโ theorem, there exists a unique selfadjoint operator $`AL()`$ such that
$$B(\xi ,\eta )=B_A(\xi ,\eta )=A\xi ,\eta ;$$
we search to characterize the set
$$๐ซ(A,๐ฎ)=๐ฌ_๐ฎ๐ฌ^{B_A}.$$
Observe that $`๐ซ(A,๐ฎ)`$ has a unique element if $`A`$ is a positive invertible operator, but in general it can have 0, 1 or infinite elements. Even if we get a characterization of $`๐ซ(A,๐ฎ)`$ in general, much more satisfactory results can be obtained for a positive $`A`$ ($`A0`$, i.e. $`A\xi ,\xi 0`$ for all $`\xi `$). In this paper, a pair $`(A,๐ฎ)`$ consisting of a bounded selfadjoint operator $`A`$ and a closed subspace $`๐ฎ`$ is said to be $`compatible`$ if $`๐ซ(A,๐ฎ)`$ is not empty.
The contents of the paper are the following:
In section 2 we collect several known results we shall use later. We show in this section that if $`Q๐ฌ`$, $`AL(H)`$ and $`R(QA)R(A)`$. Then the unique operator $`DL()`$ verifying that
$$QA=AD,\mathrm{ker}D=\mathrm{ker}QAandR(D)\overline{R(A^{})},$$
(called the reduced solution of $`AX=QA`$) satisfies also that $`D^2=D`$, i.e., $`D๐ฌ`$.
In section 3, some characterizations of the compatibility of $`(A,๐ฎ)`$ are given; some of them hold for general selfadjoint operators $`A`$, and others hold only for positive operators $`A`$. Among other properties, it is shown that an oblique projection $`Q`$ is $`A`$\- seladjoint (if $`A0`$) if and only if $`0Q^{}AQA`$ (see Lemma 3.2). We establish, also for $`A0`$, that $`๐ซ(A,๐ฎ)`$ is an affine manifold and we give a parametrization for it. When $`(A,๐ฎ)`$ is compatible, a distinguished element $`P_{A,๐ฎ}๐ซ(A,๐ฎ)`$ can be defined. It is shown that the norm of $`P_{A,๐ฎ}`$ is minimal in $`๐ซ(A,๐ฎ)`$ (see Theorem 3.6).
In section 4 we consider the relationship between the compatibility of $`(A,๐ฎ)`$ and some properties of the Schur complement. M. G. Krein and W. N. Anderson and G. E. Trapp , extended the notion of Schur complement of matrices to Hilbert space operators, defining what it is called the shorted operator. We recall the definition: if $`AL()^+`$, $`๐ฎ`$ is a closed subspace and $`P=P_๐ฎ`$ is the orthogonal projection onto $`๐ฎ`$, then the set
$$\{XL()^+:XA\text{ and }R(X)๐ฎ^{}\}$$
has a maximum (for the natural order relation in $`L()^+`$), which is called the shorted operator of $`A`$ to $`๐ฎ^{}`$. We shall denote it by $`\mathrm{\Sigma }(P,A)`$. It is shown that, for any $`Q๐ซ(A,๐ฎ)`$, the Schur complement $`\mathrm{\Sigma }(P,A)`$ verifies that
$$\mathrm{\Sigma }(P,A)=A(1Q)$$
(see Proposition 4.2). We also show that $`(A,๐ฎ)`$ is compatible if and only if, in the characterization
$$\mathrm{\Sigma }(P,A)=inf\{R^{}AR:R๐ฌ,\mathrm{ker}R=๐ฎ\},$$
due by Anderson and Trapp , the infimum is, indeed, a minimum (see Corollary 4.3).
In section 5 we consider the case of positive operators $`A`$ which are $`injective`$. Using properties of the shorted operator $`\mathrm{\Sigma }(P,A)`$, new conditions equivalent to the fact that the pair $`(A,๐ฎ)`$ is compatible are found. For example (see Proposition 5.5), it is shown that
$$(A,๐ฎ)\text{ is compatible }๐ฎ^{}R(A+\lambda (1P)),\text{ for some }\lambda >0.$$
In section 6 we consider the case of positive operators $`A`$ with $`closed`$ $`range`$. Among other equivalences, it is shown that $`(A,๐ฎ)`$ is compatible if and only if $`๐ฎ+\mathrm{ker}A`$ is closed (see Theorem 6.2). As a consequence it is shown that all manifolds $`๐ซ(B,๐ฎ)`$ for $`R(B)=R(A)`$ are โparallelโ (see Corollary 6.4). So, in this sense, it suffices to study the case of the orthogonal projection $`Q=P_{R(A)}`$. This case is studied in section 7, where we show a formula for the norm of the projection $`P_{Q,P}:=P_{Q,๐ฎ}`$ in $`๐ซ(Q,๐ฎ)`$. For example (see Proposition 7.2), if $`\mathrm{ker}QR(P)=\{0\}`$, then $`PQPGL(๐ฎ)`$ and
$$P_{Q,P}^2=(PQP)^1=(1(1Q)P^2)^1.$$
In case that $`R(P)\mathrm{ker}Q=\{0\}=R(Q)\mathrm{ker}P`$ (e.g., if $`P`$ and $`Q`$ are in position p or generic position ), $`P_{Q,P}`$ is the oblique projection given by
$$\mathrm{ker}P_{Q,P}=\mathrm{ker}Q\text{ and }R(P_{Q,P})=R(P).$$
## 2 Preliminaries
In this paper $``$ denotes a Hilbert space, $`L()`$ is the algebra of all linear bounded operators on $``$, $`L()^+`$ is the subset of $`L()`$ of all (selfadjoint) positive operators, $`GL()`$ is the group of all invertible operators in $`L()`$ and $`GL()^+=GL()L()^+`$ (positive invertible operators). For every $`CL()`$ its range is denoted by $`R(C)`$.
Denote by $`๐ฌ`$ (resp. $`๐ซ`$) the set of all projections (resp. selfadjoint projections) in $`L()`$:
$$๐ฌ=๐ฌ(L())=\{QL():Q^2=Q\},๐ซ=๐ซ(L())=\{P๐ฌ:P=P^{}\}.$$
The nonselfadjoint elements of $`๐ฌ`$ will be called oblique projections.
Along this note we use the fact that every $`P๐ซ`$ induces a representation of elements of $`L()`$ by $`2\times 2`$ matrices. Under this representation $`P`$ can be identified with
$$\left(\begin{array}{cc}I_{P()}& 0\\ 0& 0\end{array}\right)=\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)$$
and all idempotents $`Q`$ with the same range as $`P`$ have the form
$$Q=\left(\begin{array}{cc}1& x\\ 0& 0\end{array}\right)$$
for some $`xL(\mathrm{ker}P,R(P))`$.
Now we state the well known criterium due to Douglas (see also Fillmore-Williams ) about ranges and factorizations of operators:
###### Theorem 2.1
Let $`A,BL()`$. Then the following conditions are equivalent:
1. $`R(B)R(A)`$.
2. There exists a positive number $`\lambda `$ such that $`BB^{}\lambda AA^{}`$.
3. There exists $`DL()`$ such that $`B=AD`$.
Moreover, the operator $`D`$ is unique if it satisfies the conditions
$$B=AD,\mathrm{ker}D=\mathrm{ker}BandR(D)\overline{R(A^{})}.$$
In this case $`D^2=inf\{\lambda :BB^{}\lambda AA^{}\}`$ and $`A`$ is called the reduced solution of the equation $`AX=B`$.
###### Corollary 2.2
Suppose that $`Q๐ฌ`$, $`AL(H)`$ and $`R(QA)R(A)`$. Then the reduced solution $`DL()`$ of $`AX=QA`$ satisfies that $`D^2=D`$, i.e., $`D๐ฌ`$.
Proof. Note that $`AD^2=QAD=Q^2A=QA`$. Also
$$\mathrm{ker}QA=\mathrm{ker}D\mathrm{ker}D^2\mathrm{ker}AD^2=\mathrm{ker}QA$$
and $`R(D^2)R(D)\overline{R(A^{})}`$. Thus, $`D^2`$ is a reduced solution of $`AX=QA`$ and, by uniqueness, it must be $`D^2=D`$, i.e. $`D๐ฌ`$
## 3 $`A`$-selfadjoint projections, generic properties
Throughout, $`๐ฎ`$ is a closed subspace of $``$ and $`P`$ is the orthogonal projection onto $`๐ฎ`$. As we said in the introduction, we consider a bounded sesquilinear form $`B=B_A:\times \text{C}\text{ }\text{ }`$ determined by a Hermitian operator $`AL()`$:
$$B_A(\xi ,\eta )=A\xi ,\eta ,\xi ,\eta .$$
This form induces the notion of $`A`$-orthogonality. For example, easy computations show that the $`A`$-orthogonal of $`๐ฎ`$ is
$$๐ฎ^_A:=\{\xi :A\xi ,\eta =0\eta ๐ฎ\}=A^1(๐ฎ^{}).$$
Given $`TL()`$, an operator $`WL()`$ is called an A-adjoint of $`T`$ if
$$B_A(T\xi ,\eta )=B_A(\xi ,W\eta ),\xi ,\eta ,$$
or, which is the same, if
$$T^{}A=AW.$$
Observe that $`T`$ may have no $`A`$-adjoint, only one or many of them. We shall not deal in this paper with the general problem of existence and uniqueness of $`A`$-adjoint operators. Instead, we shall study the existence and uniqueness of $`A`$-selfadjoint projections, i.e., $`Q๐ฌ`$ such that $`AQ=Q^{}A`$. Among them, we are interested in those whose range is exactly $`๐ฎ`$. Thus, the main goal of the paper is the study of the set
$$๐ซ(A,๐ฎ)=\{Q๐ฌ:R(Q)=๐ฎ,AQ=Q^{}A\}$$
for different choices of $`A`$.
###### Definition 3.1
Let $`A=A^{}L()`$ and $`๐ฎ`$ a closed subspace. The pair $`(A,๐ฎ)`$ is said to be $`compatible`$ if there exists an $`A`$-selfadjoint projection with range $`๐ฎ`$, i.e. if $`๐ซ(A,๐ฎ)`$ is not empty.
For general results on $`A`$-selfadjoint operators the reader is referred to the papers by P. Lax and J. Dieudonnรฉ ; a recent paper by S.Hassi and K. Nordstrรถm contains many interesting results on $`A`$-selfadjoint projections. Some of the results of this section overlapp with their work, but we include them because the methods used in our proofs are useful for the study of the case of a positive $`A`$, which is our main concern.
###### Lemma 3.2
Let $`A=A^{}L()`$ and $`Q๐ฌ`$. Then the following conditions are equivalent:
1. $`Q`$ verifies that $`AQ=Q^{}A`$, i.e. $`Q`$ is $`A`$-selfadjoint.
2. $`\mathrm{ker}QA^1(R(Q)^{})=R(Q)^_A`$.
If $`AL()^+`$, they are equivalent to
1. $`Q^{}AQA`$.
Proof.
1 $``$ 2: If $`Q๐ซ(A,๐ฎ)`$ and $`\xi ,\eta `$, then
$$A\eta ,Q\xi =Q^{}A\eta ,\xi =AQ\eta ,\xi =Q\eta ,A\xi ,$$
(1)
so $`\mathrm{ker}QA^1(S^{})`$. The converse can be proved in a similar way.
1 $``$ 3: Suppose that $`0Q^{}AQA`$. Then, by Theorem 2.1, the reduced solution $`D`$ of the equation $`A^{1/2}X=Q^{}A^{1/2}`$ satisfies $`D1`$ and, by Corollary 2.2, $`D^2=D`$. Thus, it must be $`D^{}=D`$. Since $`Q^{}A=A^{1/2}DA^{1/2}`$, we conclude that $`Q^{}A=AQ`$. Conversely, note that $`AQ=Q^{}AQ0`$ and, if $`E=1Q`$, $`AE=E^{}AE`$. Then $`AQA`$, because, for $`\xi `$,
$$\begin{array}{cc}\hfill AQ\xi ,\xi & =AQ\xi ,Q\xi \hfill \\ & \\ & AQ\xi ,Q\xi +AE\xi ,E\xi \hfill \\ & \\ & =AQ\xi ,\xi +AE\xi ,\xi =A(Q+E)\xi ,\xi \hfill \\ & \\ & =A\xi ,\xi \text{ }\hfill \end{array}$$
Throughout, we use the matrix representation determined by $`P`$.
###### Proposition 3.3
Given $`A=A^{}L()`$, the following conditions are equivalent:
1. The pair $`(A,๐ฎ)`$ is compatible (i.e. $`๐ซ(A,๐ฎ)`$ is not empty).
2. $`R(PA)=R(PAP)`$.
3. If $`A=\left(\begin{array}{cc}a& b\\ b^{}& c\end{array}\right)`$ then $`R(b)R(a)`$.
4. $`๐ฎ+A^1(๐ฎ^{})=`$.
Proof. Note that
$$PA=\left(\begin{array}{cc}a& b\\ 0& 0\end{array}\right)\text{ and }PAP=\left(\begin{array}{cc}a& 0\\ 0& 0\end{array}\right),$$
so $`R(a)=R(PAP)R(PA)=R(a)+R(b)`$ and items 2 and 3 are equivalent. On the other hand, for any $`Q๐ฌ`$ it holds $`R(Q)=๐ฎ`$ if and only if
$$Q=\left(\begin{array}{cc}1& x\\ 0& 0\end{array}\right).$$
Easy computations show that $`Q^{}A=AQ`$ if and only if $`ax=b`$, so items 1 and 3 are equivalent by Theorem 2.1. Finally, if $`Q๐ซ(A,๐ฎ)`$ then, by Lemma 3.2, $`\mathrm{ker}QA^1(๐ฎ^{})`$, which implies 4. Conversely, if $`๐ฎ+A^1(๐ฎ^{})=`$, and if $`๐ฉ`$ is defined by $`๐ฉ=๐ฎA^1(๐ฎ^{})`$, then $`๐ฎ(A^1(๐ฎ^{})๐ฉ)=`$. The projection $`Q`$ defined by this decomposition of $``$ verifies, again by Lemma 3.2, that $`Q^{}A=AQ`$
###### Remark 3.4
1. As mentioned before, there exist operators $`TL()`$ which do not admit $`A`$-adjoint. In fact, the existence of an $`A`$-adjoint $`W`$ of $`T`$ is equivalent to the existence of a solution of the equation $`AW=T^{}A`$ and this is equivalent to $`R(T^{}A)R(A)`$. If $`Q๐ฌ`$, then the existence of an $`A`$-adjoint of $`Q`$ is also equivalent to $`R(A)=R(A)\mathrm{ker}Q^{}+R(A)R(Q)^{}`$.
2. We conjecture that the existence of some $`Q๐ฌ_๐ฎ`$ which admits $`A`$-adjoint is equivalent to the fact that $`(A,๐ฎ)`$ is compatible.
###### Definition 3.5
Let $`A=A^{}L()`$ and suppose that the pair $`(A,๐ฎ)`$ is compatible. If $`A=\left(\begin{array}{cc}a& b\\ b^{}& c\end{array}\right)`$ and $`dL(๐ฎ^{},๐ฎ)`$ is the reduced solution of the equation $`ax=b`$, we define the following oblique projection onto $`๐ฎ`$:
$$P_{A,๐ฎ}:=\left(\begin{array}{cc}1& d\\ 0& 0\end{array}\right)$$
###### Theorem 3.6
Let $`A=A^{}L()`$ and suppose that $`(A,๐ฎ)`$ is compatible. Then the following properties hold:
1. $`P_{A,๐ฎ}๐ซ(A,๐ฎ)`$.
2. $`๐ซ(A,๐ฎ)`$ has a unique element (namely, $`P_{A,๐ฎ}`$) if and only if $`๐ฎA^1(๐ฎ^{})=`$.
If $`AL()^+`$, then
1. $`A^1(S^{})๐ฎ=\mathrm{ker}A๐ฎ:=๐ฉ`$
2. $`๐ซ(A,๐ฎ)`$ is an affine manifold and it can be parametrized as
$$๐ซ(A,๐ฎ)=P_{A,๐ฎ}+L(๐ฎ^{},๐ฉ),$$
where $`L(๐ฎ^{},๐ฉ)`$ is viewed as a subspace of $`L()`$. A matrix representation of this parametrization is
$$๐ซ(A,๐ฎ)Q=P_{A,๐ฎ}+z=\left(\begin{array}{ccc}1& 0& d\\ 0& 1& z\\ 0& 0& 0\end{array}\right)\begin{array}{c}๐ฎ๐ฉ\hfill \\ ๐ฉ\hfill \\ ๐ฎ^{}\hfill \end{array}$$
(2)
with the notations of Definition 3.5.
3. $`P_{A,๐ฎ}`$ has minimal norm in $`๐ซ(A,๐ฎ)`$:
$$P_{A,๐ฎ}=\mathrm{min}\{Q:Q๐ซ(A,๐ฎ)\}.$$
Nevertheless, $`P_{A,๐ฎ}`$ is not in general the unique $`Q๐ซ(A,๐ฎ)`$ that realizes the minimum norm.
Proof.
1. Use the same argument as in the proof of Proposition 3.3.
2. By Lemma (3.2), if $`Q๐ฌ`$ and $`R(Q)=๐ฎ`$, then $`Q๐ซ(A,๐ฎ)`$ if and only if $`\mathrm{ker}QA^1(S^{})`$. This clearly implies 2.
3. If $`\xi A^1(S^{})๐ฎ`$, then $`A^{1/2}\xi =A\xi ,\xi =0`$, so that $`A\xi =0`$.
4. We have to show that every element $`Q๐ซ(A,๐ฎ)`$ can be written in an unique form as
$$Q=P_{A,๐ฎ}+z,\text{ with }zL(๐ฎ^{},๐ฉ).$$
If $`A=\left(\begin{array}{cc}a& b\\ b^{}& c\end{array}\right)`$, $`Q=\left(\begin{array}{cc}1& y\\ 0& 0\end{array}\right)`$ with $`yL(๐ฎ^{},๐ฎ)`$ and $`dL(๐ฎ^{},๐ฎ)`$ is the reduced solution of the equation $`ax=b`$, then $`Q๐ซ(A,๐ฎ)`$ if and only if $`ay=b`$ if and only if $`a(yd)=0`$. Therefore, if $`z=ydL(๐ฎ^{},๐ฎ)`$, then $`Q๐ซ(A,๐ฎ)`$ if and only if $`Q=P_{A,๐ฎ}+z`$ and $`R(z)\mathrm{ker}a`$. But
$$\mathrm{ker}a=๐ฎ\mathrm{ker}PAP=๐ฎ\mathrm{ker}A=๐ฉ.$$
Concerning the matrix representation, note that, by Theorem 2.1,
$$R(d)\overline{R(a)}=(\mathrm{ker}a)^{}=๐ฎ๐ฉ.$$
5. If $`Q๐ซ(A,๐ฎ)`$ has the matrix form given in equation (2), then
$$Q^2=1+\left(\begin{array}{ccc}0& 0& d\\ 0& 0& z\\ 0& 0& 0\end{array}\right)^21+d^2=P_{A,๐ฎ}^2.$$
Choose $`dL(๐ฎ^{},๐ฎ)`$ such that $`d=1`$, $`R(d)=\overline{R(d)}๐ฎ`$ and $`\mathrm{ker}d\{0\}`$. Then the matrix
$$A=\left(\begin{array}{cc}P_{R(d)}& d\\ d^{}& 1\end{array}\right)0,$$
$`๐ฉ=\mathrm{ker}A๐ฎ=๐ฎR(d)`$ and $`d`$ is the reduced solution of $`P_{R(d)}x=d`$. Let $`zL(\mathrm{ker}d,๐ฉ)`$ with $`0<z1`$; then the projection $`Q=P_{A,๐ฎ}+z`$ as in equation (2) satisfies $`Q๐ซ(A,๐ฎ)`$, $`Q=P_{A,๐ฎ}=\sqrt{2}`$ and $`QP_{A,๐ฎ}`$
## 4 Schur complements and $`A`$-selfadjoint projections
As before, let $`P๐ซ`$ be the orthogonal projection onto the closed subspace $`๐ฎ`$. Every $`AGL()^+`$ defines a scalar product on $``$ which is equivalent to $`,`$, namely
$$\xi ,\eta _A=A\xi ,\eta ,\xi ,\eta .$$
The unique projection $`P_{A,๐ฎ}`$ onto $`๐ฎ`$ which is $`A`$-orthogonal, i.e., $`A`$-selfadjoint, is uniquely determined by
$$P_{A,๐ฎ}=P(1+PA^1PA)^1=P(PAP+(1P)A(1P))^1A.$$
Observe that $`P_{A,๐ฎ}=A^1P_{A,๐ฎ}^{}A`$, because $`A`$ is invertible. In particular, in this case the set $`๐ซ(A,๐ฎ)`$ is a singleton. Analogously, there exists a unique projection $`Q_{A,๐ฎ}`$ which is $`A`$-orthogonal and has kernel $`๐ฎ`$: $`Q_{A,๐ฎ}=1P_{A,๐ฎ}`$. Notice that $`AQ_{A,๐ฎ}=Q_{A,๐ฎ}^{}A`$.
Consider the map
$$\mathrm{\Sigma }:๐ซ\times GL()^+L()^+,\text{ defined by }\mathrm{\Sigma }(P,A)=AQ_{A,๐ฎ}=Q_{A,๐ฎ}^{}A.$$
If $`AGL()^+`$ has matrix representation $`A=\left(\begin{array}{cc}a& b\\ b^{}& c\end{array}\right),`$ then
$$P_{A,๐ฎ}=\left(\begin{array}{cc}1& a^1b\\ 0& 0\end{array}\right),Q_{A,๐ฎ}=\left(\begin{array}{cc}0& a^1b\\ 0& 1\end{array}\right),\text{ and }\mathrm{\Sigma }(P,A)=\left(\begin{array}{cc}0& 0\\ 0& cb^{}a^1b\end{array}\right).$$
This reminds us the Schur complement. Recall that, given a square matrix $`M=\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)`$, with $`a`$ and $`d`$ square blocks, a Schur complement of $`a`$ in $`M`$ is $`dca^{}b`$, where $`a^{}`$ is a generalized inverse of $`a`$. The reader is referred to and for concise surveys on the subject. This notion has been extended to positive Hilbert space operators by M. G. Krein and, later and independently, by W. N. Anderson and G. E. Trapp defining what is called the shorted operator: if $`AL()^+`$ then the set
$$\{XL()^+:XA\text{ and }R(X)๐ฎ^{}\}$$
has a maximum (for the natural order relation in $`L()^+`$), which is called the shorted operator of $`A`$ to $`๐ฎ^{}`$.
Next we collect some results of Anderson-Trapp and E. L. Pekarev which are relevant in this paper. Observe that the first item allows us to extend the map $`\mathrm{\Sigma }`$ to $`๐ซ\times L()^+.`$
###### Theorem 4.1
Let $`AL()^+`$ with matrix representation
$$A=\left(\begin{array}{cc}a& b\\ b^{}& c\end{array}\right).$$
1. If $`A`$ is invertible, then $`\mathrm{\Sigma }(P,A)`$ coincides with the shorted operator of $`A`$ to $`๐ฎ^{}`$. We shall keep the notation $`\mathrm{\Sigma }(P,A)`$ for the shorted operator of $`A`$ to $`R(P)^{}`$ for every pair $`(P,A)๐ซ\times L()^+`$.
2. $`R(b)R(a^{1/2})`$ and if $`dL()`$ is the reduced solution of the equation $`a^{1/2}x=b`$ then
$$\mathrm{\Sigma }(P,A)=\left(\begin{array}{cc}0& 0\\ 0& cd^{}d\end{array}\right)$$
3. If $`=A^{1/2}(๐ฎ^{})`$ and $`P_{}`$ is the orthogonal projection onto $``$ then
$$\mathrm{\Sigma }(P,A)=A^{1/2}P_{}A^{1/2}.$$
4. $`\mathrm{\Sigma }(P,A)`$ is the infimum of the set $`\{R^{}AR:R๐ฌ,\mathrm{ker}R=๐ฎ\}`$; in general, the infimum is not attained.
5. $`R(A)๐ฎ^{}R(\mathrm{\Sigma }(P,A))R(\mathrm{\Sigma }(P,A)^{1/2})=R(A^{1/2})๐ฎ^{}`$; in general, the inclusions are strict.
The reader is referred to and for proofs of these facts. We prove now that the infimum of item 4 is attained if and only if $`(A,๐ฎ)`$ is compatible by relating the notions of shorted operators and $`A`$-selfadjoint projections (when there is one). As a consequence, we complete item 5 of the last theorem in case that $`(A,๐ฎ)`$ is compatible.
###### Proposition 4.2
Let $`AL()^+`$ such that the pair $`(A,๐ฎ)`$ is compatible. Let $`E๐ซ(A,๐ฎ)`$ and $`Q=1E`$. Then
1. $`\mathrm{\Sigma }(P,A)=AQ=Q^{}AQ`$.
2. $`\mathrm{\Sigma }(P,A)=\mathrm{min}\{R^{}AR:R๐ฌ,\mathrm{ker}R=๐ฎ\}`$ and the minimum is attained at $`Q`$.
3. $`R(\mathrm{\Sigma }(P,A))=R(A)S^{}`$.
Proof.
1. Note that $`0AQ=Q^{}AQA`$, by Lemma 3.2. Also $`R(AQ)=R(Q^{}A)R(Q^{})=๐ฎ^{}`$. Given $`XA`$ with $`R(X)๐ฎ^{}`$, then, since $`\mathrm{ker}Q=๐ฎ`$, we have that
$$X=Q^{}XQQ^{}AQ=AQ,$$
where the first equality can be easily checked because $`X`$ has the form $`\left(\begin{array}{cc}0& 0\\ 0& x\end{array}\right)`$.
2. By item 1, $`Q^{}AQ=\mathrm{\Sigma }(P,A)`$ and $`\mathrm{ker}Q=๐ฎ`$. So the minimum is attained at $`Q`$ by Theorem 4.1.
3. Clearly the equation $`\mathrm{\Sigma }(P,A)=AQ`$ implies that $`R(\mathrm{\Sigma }(P,A))R(A)S^{}`$. The other inclusion always holds by Theorem 4.1
###### Corollary 4.3
If $`AL()^+`$ the following conditions are equivalent:
1. The pair $`(A,๐ฎ)`$ is compatible.
2. The set $`\{S^{}AS:S๐ฌ,\mathrm{ker}S=๐ฎ\}`$ attains its minimum at some projection $`R`$.
3. There exists $`R๐ฌ`$ such that $`\mathrm{ker}R=๐ฎ`$ and $`R^{}ARA`$.
Proof. 1 $``$ 2: Follows from Proposition 4.2.
2 $``$ 3: Follows from Theorem 4.1.
3 $``$ 1: By Lemma 3.2, any projection $`R`$ such that $`R^{}ARA`$ verifies that $`AR=R^{}A`$. If also $`\mathrm{ker}R=๐ฎ`$, then $`1R๐ซ(A,๐ฎ)`$
In the next sections we shall study the existence of $`A`$-selfadjoint projections onto a closed subspace $`๐ฎ`$, under particular hypothesis on the positive operator $`A`$.
## 5 $`A`$-selfadjoint projections: the injective case
As before, let $`P๐ซ`$ be the orthogonal projection onto $`๐ฎ`$. In this section we study the case of injective operators $`AL()^+`$. We define the notion of $`A`$-admissibility for $`๐ฎ`$, in terms of the shorted operator $`\mathrm{\Sigma }(P,A)`$. This notion is shown to be strictly weaker that compatibility for the pair $`(A,๐ฎ)`$. Under the assumption of $`A`$-admissibitity for $`๐ฎ`$, the fact that $`(A,๐ฎ)`$ is compatible becomes equivalent to the equality $`R(\mathrm{\Sigma }(P,A))=๐ฎ^{}R(A)`$ (see item 5 of Theorem 4.1 and item 3 of Proposition 4.2).
###### Lemma 5.1
Given $`AL()^+`$ which is injective and $`=A^{1/2}(๐ฎ^{})`$, the following conditions are equivalent:
1. $`\mathrm{ker}\mathrm{\Sigma }(P,A)=๐ฎ`$
2. $`^{}A^{1/2}(๐ฎ^{})=\{0\}`$.
3. $`\overline{๐ฎ^{}R(A^{1/2})}=๐ฎ^{}`$
4. $`๐ฎ=(๐ฎ^{}R(A^{1/2}))^{}`$
5. $`๐ฎ=๐ฏ^{}`$ for some subspace $`๐ฏR(A^{1/2})`$
Proof. 1 $``$ 2: Recall that $`\mathrm{\Sigma }(P,A)=A^{1/2}P_{}A^{1/2}`$ by Theorem 4.1. Using that $`A^{1/2}`$ is injective, we deduce 2.
2 $``$ 3: Suppose that 3 is false. Let $`\xi ๐ฎ^{}\overline{๐ฎ^{}R(A^{1/2})}`$, $`\xi 0`$. If $`\eta `$ then $`A^{1/2}\eta ๐ฎ^{}R(A^{1/2})`$ and
$$A^{1/2}\xi ,\eta =\xi ,A^{1/2}\eta =0.$$
Thus, $`A^{1/2}\xi ^{}A^{1/2}(๐ฎ^{})`$ and $`A^{1/2}\xi 0`$, which contradicts 2.
3 $``$ 4 $``$ 5: It is clear.
2 $``$ 3: If $`๐ฎ=๐ฏ^{}`$ with $`๐ฏR(A^{1/2})`$, then $`๐ฎ^{}=\overline{๐ฏ}`$ and
$$๐ฏR(A^{1/2})๐ฎ^{}=R(\mathrm{\Sigma }(P,A)^{1/2}).$$
So $`R(\mathrm{\Sigma }(P,A)^{1/2})`$ is dense in $`๐ฎ^{}`$ and $`\mathrm{ker}\mathrm{\Sigma }(P,A)=\mathrm{ker}\mathrm{\Sigma }(P,A)^{1/2}=R(\mathrm{\Sigma }(P,A)^{1/2})^{}=๐ฎ`$
###### Definition 5.2
We shall say that $`๐ฎ`$ is $`A`$-admissible if any of the conditions of Lemma 5.1 is verified.
###### Lemma 5.3
If $`AL()^+`$is injective and $`(A,๐ฎ)`$ is compatible, then $`๐ฎ`$ is $`A`$-admissible.
Proof. Let $`E๐ซ(A,๐ฎ)`$ and $`Q=1E`$. Then, by Proposition 4.2, $`\mathrm{\Sigma }(P,A)=AQ`$ and $`\mathrm{ker}\mathrm{\Sigma }(P,A)=\mathrm{ker}Q=๐ฎ`$
###### Remark 5.4
If $`(A,๐ฎ)`$ is compatible then a condition which is stronger than $`A`$-admissibility is verified. Indeed, $`\mathrm{\Sigma }(P,A)=A(1P_{A,๐ฎ})`$ implies that $`\mathrm{ker}\mathrm{\Sigma }(P,A)=๐ฎ`$. But in this case, $`R(\mathrm{\Sigma }(P,A))R(A)๐ฎ^{}`$ which must be dense in $`๐ฎ^{}`$. Note that $`R(A^{1/2})`$ strictly contains $`R(A)`$ if $`R(A)`$ is not closed.
Nevertheless, we restrict ourselves to the weaker notion of $`A`$-admissibility because under the hypothesis that $`๐ฎ`$ is $`A`$-admissible, the conditions $`(A,๐ฎ)`$ is compatible and $`R(\mathrm{\Sigma }(P,A))R(A)`$ become equivalent. Observe that $`R(\mathrm{\Sigma }(P,A))R(A)`$ is false in general (recall item 5 of Theorem 4.1 and item 3 of Proposition 4.2)
###### Proposition 5.5
If $`AL()^+`$ is injective then the following conditions are equivalent:
1. The pair $`(A,๐ฎ)`$ is compatible.
2. 1. $`\mathrm{ker}\mathrm{\Sigma }(P,A)=๐ฎ`$ (i.e. $`๐ฎ`$ is $`A`$-admissible) and
2. $`R(\mathrm{\Sigma }(P,A))R(A)`$.
3. $`๐ฎ`$ is $`A`$-admissible and, if $`=A^{1/2}(๐ฎ^{})`$, then $`P_{}AP_{}\mu A`$ for some $`\mu >0`$.
4. $`๐ฎ^{}R(A+\lambda (1P))`$ for some (and then for any) $`\lambda >0`$.
Proof.
1 $``$ 2: By Lemma 5.3, $`๐ฎ`$ must be $`A`$-admissible. If $`Q_{A,๐ฎ}=1P_{A,๐ฎ}`$, then $`\mathrm{\Sigma }(P,A)=AQ_{A,๐ฎ}`$ and 2 follows.
2 $``$ 3: If $`R(A^{1/2}P_MA^{1/2})R(A)`$ then $`R(P_MA^{1/2})R(A^{1/2})`$ and 3 holds.
3 $``$ 1: Note that $`P_MAP_M\mu A`$ if and only if $`R(P_MA^{1/2})R(A^{1/2})`$ if and only if there exists a unique $`FL()`$ such that $`A^{1/2}F=P_MA^{1/2}`$, $`\mathrm{ker}(P_MA^{1/2})\mathrm{ker}F`$ and $`R(F)R(A^{1/2})`$. We shall see that $`1F๐ซ(A,๐ฎ)`$. Indeed, $`F^2=F`$ by Corollary 2.2. $`F`$ is $`A`$-selfadjoint because $`AF=A^{1/2}P_MA^{1/2}=\mathrm{\Sigma }(P,A)`$ which is selfadjoint. Finally, $`\mathrm{ker}F=๐ฎ`$. Indeed, $`AF=\mathrm{\Sigma }(P,A)`$, so $`\mathrm{ker}F=\mathrm{ker}\mathrm{\Sigma }(P,A)=๐ฎ`$ because $`๐ฎ`$ is $`A`$-admissible.
4 $``$ 1: Using Proposition 3.3, we know that the fact that $`(A,๐ฎ)`$ is compatible only depends on the first row $`PA`$ of $`A`$. Therefore we can freely change $`A`$ by $`A+\lambda (1P)`$, for $`\lambda >0`$. In this case conditions 2 can be rewritten as condition 4, since $`\mathrm{\Sigma }(P,A+\lambda (1P))=\mathrm{\Sigma }(P,A)+\lambda (1P)`$.
###### Example 5.6
Given a positive injective operator $`AL()`$ with non-closed range, it is easy to show that there exists $`\xi R(A^{1/2})R(A)`$. Let $`P_\xi `$ be the orthogonal projection onto the subspace generated by $`\xi `$. Then $`R(P_\xi )R(A^{1/2})`$, so that, by Douglasโ theorem, $`P_\xi \lambda A`$ for some positive number $`\lambda `$ which we can suppose equal to $`1`$, by changing $`A`$ by $`\lambda A`$. It is well known that this implies that the operator $`BL()`$ defined by
$$B=\left(\begin{array}{cc}A& P_\xi \\ P_\xi & A\end{array}\right)$$
is positive. Let $`๐ฎ=_1=0`$. Then $`๐ฎ^{}=_2=0`$. We shall see that $`B`$ is injective, $`_1`$ is $`B`$-admissible, moreover $`_2R(B)`$ is dense in $`_2`$, but $`๐ซ(B,๐ฎ)`$ is empty.
Indeed, it is clear that $`B`$ does not verify condition 3 of Proposition 3.3, so $`๐ซ(B,๐ฎ)`$ is empty. Let $`D`$ be the reduced solution of $`P_\xi =A^{1/2}X`$. Then $`\mathrm{\Sigma }(P,B)=AD^{}D`$. Note that $`\mathrm{ker}D=\mathrm{ker}P_x`$ implies $`DP_\xi =D`$. So $`D^{}D=P_\xi D^{}D`$. Then, if $`0\eta \mathrm{ker}\mathrm{\Sigma }(P,B)`$,
$$A\eta =D^{}D\eta =P_\xi D^{}D\eta =\lambda \xi \text{ for some }\lambda \text{RI }\eta =0$$
because $`\xi R(A)`$ and $`A`$ is injective. So $`\mathrm{ker}\mathrm{\Sigma }(P,B)=๐ฎ`$ and $`_1`$ is $`B`$-admissible. Also
$$B(\omega \eta )0A\omega +P_\xi \eta =0\omega =0\text{ and }\eta \{\xi \}^{}.$$
(3)
Then $`R(B)_2=\{B(0\eta ):\eta \{\xi \}^{}\}=0A(\{\xi \}^{})`$. We shall see that $`A(\{\xi \}^{})`$ is dense in $``$. Indeed, if $`\zeta [A(\{\xi \}^{})]^{}`$, then $`\eta ,A\zeta =A\eta ,\zeta =0`$ for all $`\eta \{\xi \}^{}`$. So $`A\zeta =\mu \xi `$ for some $`\mu \text{RI }`$. As before this implies that $`\zeta =0`$. Finally, the injectivity of $`B`$ can be easily deduced from equation (3).
## 6 $`A`$-selfadjoint projections: the closed range case
As before we fix $`P๐ซ`$ with $`R(P)=๐ฎ`$. In this section $`A`$ denotes a positive operator with closed range. We shall see that, in this case, the fact that $`(A,๐ฎ)`$ is compatible depends only on the angle between $`\mathrm{ker}A`$ and $`๐ฎ`$, i. e. $`(A,๐ฎ)`$ is compatible if and only if $`\mathrm{ker}A+๐ฎ`$ is closed. To establish the link between compatibility and the angle condition, we need to determine when $`R(PAP)`$ is closed. This is done in the following Lemma:
###### Lemma 6.1
It holds that
$$\overline{R(PAP)}=๐ฎ(๐ฎ\mathrm{ker}A)^{}$$
and that $`R(PAP)`$ is closed if and only if the subspace $`\mathrm{ker}A+๐ฎ`$ is closed.
Proof. First note that $`\mathrm{ker}PAP=\{x:PAPx,x=0\}=\{x:Px\mathrm{ker}A\}=\mathrm{ker}AP.`$ So $`\mathrm{ker}PAP=๐ฎ^{}(๐ฎ\mathrm{ker}A)`$. Therefore
$$\overline{R(PAP)}=(\mathrm{ker}PAP)^{}=๐ฎ(๐ฎ\mathrm{ker}A)=๐ฎ(๐ฎ\mathrm{ker}A)^{}:=.$$
Clearly $`\mathrm{ker}A=\{0\}`$. Suppose that $`๐ฉ=\mathrm{ker}A+๐ฎ=\mathrm{ker}A`$ is closed. Let $`Q`$ the projection from $`๐ฉ`$ onto $``$ with $`\mathrm{ker}Q=\mathrm{ker}A`$; observe that $`Q`$ is bounded. If $`Q=0`$ then $`=\{0\}`$, $`๐ฎ\mathrm{ker}A`$ and $`PAP=0`$. If $`\{0\}`$, given $`\xi `$, let $`\eta R(A)`$ such that $`A\eta =A\xi `$ ($`A`$ is invertible in $`R(A)`$). Clearly $`\eta =\xi +\zeta `$ with $`\zeta \mathrm{ker}A`$. Then $`\eta ๐ฉ`$, $`Q\eta =\xi `$ and $`\xi Q\eta `$. Therefore
$$PAP\xi ,\xi =A\xi ,\xi =A\eta ,\eta \lambda \eta ^2\lambda Q^2\xi ^2.$$
for some $`\lambda >0`$, since $`A|_{R(A)}`$ is bounded from below. Conversely, if $`R(PAP)`$ is closed then $`R(PAP)=`$. Then there exists $`\mu >0`$ such that $`A\xi ,\xi =A^{1/2}\xi ^2\mu \xi ^2`$ for $`\xi `$ and $`A^{1/2}()`$ is closed. So $`๐ฉ=A^{1/2}(A^{1/2}())`$ must be also closed
###### Theorem 6.2
If $`AL()^+`$ has closed range then the following conditions are equivalent:
1. The pair $`(A,๐ฎ)`$ is compatible.
2. $`R(PAP)`$ is closed.
3. $`๐ฎ+\mathrm{ker}A`$ is closed.
4. $`R(PA)`$ is closed.
5. $`๐ฎ^{}+R(A)`$ is closed.
6. $`R(AP)=A(๐ฎ)`$ is closed.
Proof. By Lemma 6.1 conditions 2 and 3 are equivalent.
2 $``$ 1: Let $`A=\left(\begin{array}{cc}a& b\\ b^{}& c\end{array}\right)`$ in terms of $`P`$. Note that $`a=PAP`$, so $`R(a)`$ is closed. Therefore, since $`A0`$, $`R(b)R(a^{1/2})=R(a)`$. Then $`(A,๐ฎ)`$ is compatible by Proposition 3.3.
1 $``$ 3: Suppose that $`(A,๐ฎ)`$ is compatible. Let $`P_{A,๐ฎ}๐ซ(A,๐ฎ)`$ and let $`Q_{A,๐ฎ}=1P_{A,๐ฎ}`$. Then
$$\begin{array}{cc}\hfill \mathrm{ker}A& \mathrm{ker}(Q_{A,๐ฎ}^{^{}}A)=\mathrm{ker}(AQ_{A,๐ฎ})\hfill \\ & \\ & =\{\xi :Q_{A,๐ฎ}\xi \mathrm{ker}A\}\hfill \\ & \\ & =๐ฎ(\mathrm{ker}AR(Q_{A,๐ฎ}))\hfill \\ & \\ & \mathrm{ker}A+๐ฎ.\hfill \end{array}$$
Therefore $`\mathrm{ker}A+๐ฎ=(\mathrm{ker}AQ_{A,๐ฎ})`$ which is closed.
4 $``$ 5: This is an easy consequence of the identity
$$R(A)+๐ฎ^{}=P^1[P(R(A))]=P^1[R(PA)].$$
3 $``$ 5: In fact, it holds in general that the sum of two closed subspaces is closed if and only if the sum of their orthogonal complements is closed (see ).
4 $``$ 6 : It is a general fact that $`R(C)`$ is closed if and only if $`R(C^{})`$ is closed.
###### Remark 6.3
Conditions 3, 4 and 5, 6 are known to be equivalent, since $`R(P)=๐ฎ`$ and $`\mathrm{ker}P=๐ฎ^{}`$ (see Thm. 22 of ). They are also equivalent to, for example, the angle condition
$$c(๐ฎ,\mathrm{ker}A)<1,$$
where $`c(๐ฎ,๐ฏ)`$ is the cosine of the Friedrichs angle between the two subspaces $`๐ฎ,๐ฏ`$, defined by:
$$c(๐ฎ,๐ฏ)=sup\{|\xi ,\eta |:\xi ๐ฎ(๐ฎ๐ฏ)^{},\xi 1,\eta ๐ฏ(๐ฎ๐ฏ)^{},\eta 1\}$$
(4)
Also Lemma 6.1 can be deduced from the results of .
###### Corollary 6.4
For every $`AL()^+`$ with closed range, the following conditions are equivalent:
1. The pair $`(A,๐ฎ)`$ is compatible.
2. For all $`BL()^+`$ with $`R(B)=R(A)`$, the pair $`(B,๐ฎ)`$ is compatible.
3. The pair $`(P_{R(A)},๐ฎ)`$ is compatible, if $`P_{R(A)}`$ denotes the orthogonal projection onto the closed subspace $`R(A)`$
Moreover, if $`BL()^+`$ and $`R(B)=R(A)`$, then the affine manifolds $`๐ซ(A,๐ฎ)`$ and $`๐ซ(B,๐ฎ)`$ are โparallelโ, i.e.
$$๐ซ(B,๐ฎ)=(P_{B,๐ฎ}P_{A,๐ฎ})+๐ซ(A,๐ฎ).$$
(5)
Proof. If $`R(B)=R(A)`$ then $`\mathrm{ker}B=\mathrm{ker}A=\mathrm{ker}P_{R(A)}`$ and, by Theorem 6.2, the three conditions are equivalent. Equality (5) follows from the parametrization given in Theorem 3.6, since
$$A^1(๐ฎ^{})๐ฎ=\mathrm{ker}A๐ฎ=\mathrm{ker}B๐ฎ=B^1(๐ฎ^{})๐ฎ\text{ }$$
Condition 3 is an invitation to consider the sets $`๐ซ(Q,๐ฎ)`$ for $`Q๐ซ`$, which we study in the next section.
## 7 The case of two projections
In this section we shall study the case in which $`A`$ is an orthogonal projection, i.e., $`A=Q๐ซ`$. Then, by Theorem 6.2 (items 3 and 6), $`\mathrm{ker}Q+R(P)`$ is closed if and only $`๐ซ(Q,R(P))`$ is not empty. In this case we shall denote by $`P_{Q,P}`$ the projection $`P_{Q,R(P)}`$ of Definition 3.5. In the following theorem we collect several conditions which are equivalent to the existence of $`P_{Q,P}`$. Notice, however, that the equivalence of items 3 to 10 can be deduced from results by R. Bouldin and S. Izumino ; a nice survey on this and related subjects can be found in . Observe that Theorem 6.2 provides alternative proofs of some of the equivalences.
###### Theorem 7.1
Let $`P,Q๐ซ`$ with $`R(P)=๐ฎ`$ and $`R(Q)=๐ฏ`$. The following are equivalent:
1. $`(Q,๐ฎ)`$ is compatible.
2. $`(P,๐ฏ)`$ is compatible.
3. $`\mathrm{ker}Q+R(P)`$ is closed.
4. $`\mathrm{ker}P+R(Q)`$ is closed.
5. $`R(PQ)`$ is closed.
6. $`R(QP)`$ is closed.
7. $`R(1P+Q)`$ is closed.
8. $`R(1Q+P)`$ is closed.
9. $`c(๐ฎ,๐ฏ^{})=c(๐ฏ,๐ฎ^{})<1`$.
If $`\mathrm{ker}Q๐ฎ=\{0\}`$, they are equivalent to
1. $`(1Q)P<1`$.
Proof.
1 $``$ 2 $``$ 3: Follows from Theorem 6.2.
3 $``$ 4 $``$ 9: Follows from theorem 13 of .
3 $``$ 6 and 4 $``$ 5: Follows from theorem 22 of .
5 $``$ 7 and 6 $``$ 8: Follows from 2.5 of
3 $``$ 10 Follows from theorem 13 of
Suppose that any of the conditions of Theorem 7.1 is verified by $`P,Q๐ซ`$. As a final result, we shall compute $`P_{Q,P}`$. First, we assume that $`\mathrm{ker}QR(P)=\{0\}`$:
###### Proposition 7.2
Let $`P,Q๐ซ`$. Denote $`R(P)=๐ฎ`$. Suppose that $`\mathrm{ker}Q๐ฎ=\{0\}`$ and $`\mathrm{ker}Q+๐ฎ`$ is closed. Then $`Q|_๐ฎ`$ is invertible in $`L(๐ฎ,Q(๐ฎ))`$, $`PQP`$ is invertible in $`L(๐ฎ)`$ and
$$P_{Q,P}=(Q|_๐ฎ)^1=(PQP)^1^{1/2}=(1(1Q)P^2)^{1/2}$$
Proof. Using Theorem 7.1, we know that $`(1Q)P<1`$. Then
$$PPQP=P(1Q)P=(1Q)P^2<1,$$
showing that $`PQP`$ is invertible in $`L(๐ฎ)`$. On the other hand consider $`Q|_๐ฎ:๐ฎQ(๐ฎ)`$. By Theorem 6.2, $`Q(๐ฎ)`$ is closed, so $`Q|_๐ฎ`$ is invertible in $`L(๐ฎ,Q(๐ฎ))`$.
If $`P_{Q,P}=\left(\begin{array}{cc}1& d\\ 0& 0\end{array}\right)`$, then $`P_{Q,P}^2=1+d^2`$. Recall that $`d`$ is the reduced solution of the equation $`PQPX=PQ(1P)`$. So, by Theorem 2.1,
$$\begin{array}{cc}\hfill d^2& =inf\{\lambda >0:PQ(1P)QP\lambda PQPQP\}\hfill \\ & \\ & =inf\{\lambda >0:PQP(1+\lambda )(PQP)^2\}\hfill \\ & \\ & =inf\{\lambda >0:P(1+\lambda )PQP\}=inf\{\lambda >0:(PQP)^1(1+\lambda )P\}\hfill \\ & \\ & =(PQP)^11.\hfill \end{array}$$
So $`P_{Q,P}^2=(PQP)^1`$. Note also that
$$P(1+\lambda )PQP\xi ^2(1+\lambda )PQP\xi ,\xi =(1+\lambda )Q\xi ^2\text{ for all }\xi ๐ฎ.$$
Taking infimum over $`\lambda `$, we get $`P_{Q,P}=(1+d^2)^{1/2}=(Q|_๐ฎ)^1.`$
It is easy to see that, if $`0<AI`$ in $`L()`$, then $`IA=1A^1^1`$. Applying this identity to $`PQP`$ in $`L(๐ฎ)`$ we get
$$(PQP)^1=(1PPQP)^1=(1P(1Q)P)^1=(1(1Q)P^2)^1\text{ }$$
###### Remark 7.3
Let $`P,Q๐ซ`$ with $`R(P)=๐ฎ`$ and $`R(Q)=๐ฏ`$ and suppose that any of the conditions of Theorem 7.1 hold. By Proposition 3.6,
$$\begin{array}{cc}\hfill \mathrm{ker}P_{Q,P}& =Q^1(\mathrm{ker}P)(\mathrm{ker}QR(P))\hfill \\ & \\ & =(\mathrm{ker}Q+R(Q)\mathrm{ker}P)(\mathrm{ker}QR(P)).\hfill \end{array}$$
Therefore, in the case that
$$R(Q)\mathrm{ker}P=\{0\}=\mathrm{ker}QR(P)$$
(6)
(e.g., if $`P`$ and $`Q`$ are in position p or generic position ) we can conclude that $`P_{Q,P}`$ is the projection given by
$$\mathrm{ker}P_{Q,P}=\mathrm{ker}Q\text{ and }R(P_{Q,P})=R(P)$$
Then $`๐ฎ\mathrm{ker}Q=`$ and $`P_{Q,P}`$ is the oblique projection given by this decomposition of $``$. In this case, formula $`P_{Q,P}=(1(1Q)P^2)^{1/2}`$ has been proved by Ptak in (see also ).
###### Theorem 7.4
Let $`P,Q๐ซ`$ which verify that $`\mathrm{ker}Q+R(P)`$ is closed. Denote by $`๐ฉ=\mathrm{ker}QR(P)`$, $`=R(P)๐ฉ`$ and $`P_0=P_{}`$. Then
1. $`๐ซ(Q,)`$ has only one element, namely $`P_{Q,P_0}`$,
2. $`P_{Q,P}=P_๐ฉ+P_{Q,P_0}`$ and
3. $`P_{Q,P}=P_{Q,P_0}=(1(1Q)P_0^2)^{1/2}`$.
Proof. If $`๐ฉ=\{0\}`$, we can use Proposition 7.2. Assume now that $`๐ฉ`$ is not trivial. Then, by the results of section 3, we get the matrix form
$$P_{Q,P}=\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& d\\ 0& 0& 0\end{array}\right)\begin{array}{c}๐ฉ\hfill \\ \hfill \\ \mathrm{ker}P\hfill \end{array}.$$
Denote
$$T=P_{Q,P}P_๐ฉ=\left(\begin{array}{ccc}0& 0& 0\\ 0& 1& d\\ 0& 0& 0\end{array}\right)\begin{array}{c}๐ฉ\hfill \\ \hfill \\ \mathrm{ker}P\hfill \end{array}.$$
(7)
We must show that $`T=P_{Q,P_0}`$. Note that $`\mathrm{ker}Q=\{0\}`$, so $`๐ซ(Q,)`$ has, at most, one element. On the other hand, $`T^2=T`$ and $`R(T)=`$ by equation (7). Also
$$T^{}Q=(T^{}+P_๐ฉ)Q=P_{Q,P}^{}Q=QP_{Q,P}=Q(P_๐ฉ+T)=QT,$$
because $`QP_๐ฉ=0`$. So, $`T=P_{Q,P_0}`$ as claimed. By equation (7) and Proposition 7.2,
$$P_{Q,P}=P_๐ฉ+P_{Q,P_0}\text{ and}$$
$$P_{Q,P}=P_{Q,P_0}=(1(1Q)P_0^2)^{1/2},$$
because $`\mathrm{ker}QR(P_0)=\{0\}`$
|
warning/0006/cond-mat0006254.html
|
ar5iv
|
text
|
# Field theory of absorbing phase transitions with a non-diffusive conserved field
## Abstract
We investigate the critical behavior of a reaction-diffusion system exhibiting a continuous absorbing-state phase transition. The reaction-diffusion system strictly conserves the total density of particles, represented as a non-diffusive conserved field, and allows an infinite number of absorbing configurations. Numerical results show that it belongs to a wide universality class that also includes stochastic sandpile models. We derive microscopically the field theory representing this universality class.
The directed percolation (DP) universality class is recognized as the canonical example of the critical behavior in the transition from an active to a single absorbing state. This universality class appears to be very robust with respect to microscopic modifications, and non-DP behavior emerges only in the presence of additional symmetries, such as in the case of symmetric absorbing states, long-range interactions, or infinitely many absorbing states.
Recently a new universality class of absorbing-state phase transitions (APT) coupled to a non-diffusive conserved field has been identified. This class characterizes the critical behavior of several models showing absorbing transitions with a dynamics that strictly conserves the density of particles, that is represented by a conserved static (non-diffusive) field. The models are tuned to criticality by varying the particle density in the initial state, and exhibit an infinite number of absorbing states. This universality class is particularly interesting because it embraces also the large group of stochastic sandpile models (and in particular the Manna model ) which are the prototypical examples that illustrate the ideas of self-organized criticality (SOC) . These are driven dissipative models in which sand (or energy) is injected into the system and dissipated through the boundaries, leading eventually to a stationary state. In the limit of infinitesimally slow external driving, the systems approach a critical state characterized by an avalanche-like response. Recently it has been pointed out that this critical state is equivalent to the absorbing phase transition present in the fixed energy case, that is, in automata that consider the same microscopic rules defining the sandpile dynamics, but without driving nor dissipation.
The numerical evidence for the existence of such a general universality class is corroborated by the observation that all the models analyzed share the same structure and basic symmetries; namely, a conserved and static non-critical field dynamically coupled to a non-conserved order parameter field, identified as the density of the active particles. These observations have led to the conjecture that, in absence of additional symmetries, all stochastic models with an infinite number of absorbing states in which the order parameter evolution is coupled to a non-diffusive conserved field define a unique universality class .
In this Letter, we study the non-diffusive field limit for the two species reaction-diffusion model introduced in Ref. (see also Ref. ). In this limit the model has a phase transition with infinitely many absorbing states and it conserves the total number of particles, that is associated to a non-diffusive conserved field. We present extensive numerical simulations of the model in two and three dimensions, and determine the full set of critical exponents. The obtained values are compatible with the new universality class conjectured in Ref.. This definitely shows the existence of a very broad universality class that includes reaction-diffusion processes, stochastic sandpile models, and lattice gases with the same symmetry properties. For the present reaction-diffusion model, it is possible to derive microscopically a field theory (FT) description. The resulting action and Langevin equations exhibit the basic symmetries that characterize this universality class, and represent the first microscopic derivation of a FT for sandpile models. Notably, the resulting FT description recovers a phenomenological Langevin approach proposed for stochastic sandpiles. The analysis provided here is a very promising path for a coherent description of several nonequilibrium critical phenomena now rationalized in a single universality class.
We consider the two-component reaction-diffusion process identified by the following set of reaction equations:
$`B`$ $``$ $`A\text{with rate }k_1.`$ (1)
$`B+A`$ $``$ $`2B\text{with rate }k_2,`$ (2)
In this system, $`B`$ particles diffuse with diffusion rate $`D_BD`$, and A particles do not diffuse, that is, $`D_A=0`$. This corresponds to the limit $`D_A0`$ of the model introduced in Ref. . From the rate Eqs. (1) and (2), it is clear that the dynamics conserves the total density of particles $`\rho =\rho _A+\rho _B`$, where $`\rho _i`$ is the density of component $`i=A,B`$. In this conserved reaction-diffusion model, the only dynamics is due to $`B`$ particles, that we can identify as active particles. $`A`$ particles do not diffuse and cannot generate spontaneously $`B`$ particles. More specifically, $`A`$ particles can only move via the motion of $`B`$ particles that later on transform into $`A`$ because of Eq. (2). In absence of $`B`$ particles, $`\rho _A`$ can be considered a static field. This implies that any configuration devoid of $`B`$ particles is an absorbing state in which the system is trapped forever.
It is easy to see that the reaction-diffusion process defined by Eqs. (1) and (2) exhibits a phase transition from an active to an absorbing phase for a non-trivial value of the total particle density $`\rho =\rho _c`$, which acts as the control parameter. The critical value $`\rho _c`$ depends upon the reaction rates $`k_1,k_2`$. The nature of this phase transition (whether it is first or second order) for $`D_A0`$ appears to be determined by the ratio between $`D_B`$ and $`D_A`$; the static field case ($`D_A=0`$), on the other hand, has never been explored. It is clear that the static field conserved reaction-diffusion (SFCRD) model allows, for any density $`\rho `$, an infinite number (in the thermodynamic limit) of absorbing configurations, in which there are no $`B`$ particles. This is the key difference with respect to the case in which $`D_A0`$, as in Ref.. In the latter case a configuration devoid of $`B`$ particles consists of many diffusing $`A`$ particles, jumping from site to site. In the long run, all particles can visit all sites, and therefore, in a statistical sense, all configuration with a fixed number of $`A`$โs are equivalent and the absorbing state can be considered statistically unique .
The present SFCRD model seems to possess all the required symmetry (stochastic dynamics, many absorbing states, static conserved field) for being part of the universality class conjectured in Ref.. In order to test this possibility, we have performed numerical simulations of the model in a $`d`$-dimensional hypercubic lattice with $`N=L^d`$ sites. Each site can store any number of $`A`$ and $`B`$ particles; that is, our model can be represented by bosonic variables. Initial conditions are generated by randomly placing $`N\rho _A^{(0)}`$ particles $`A`$ and $`N\rho _B^{(0)}`$ particles $`B`$, corresponding to a particle density $`\rho =\rho _A^{(0)}+\rho _B^{(0)}`$. The results are independent of the particular initial ratio $`\rho _A^{(0)}/\rho _B^{(0)}`$, apart from very early time transients. The dynamics proceeds in parallel. Each time step, we update the lattice according to the following rules. a) Diffusion: on each lattice site, each $`B`$ particles moves into a randomly chosen nearest neighbor site. b) After all sites have been updated for diffusion, we perform the reactions: i) On each lattice site, each $`B`$ particle is turned into an $`A`$ particle with probability $`r_1`$. ii) At the same time, each $`A`$ particle becomes a $`B`$ particle with probability $`1(1r_2)^{n_B}`$, where $`n_B`$ is the total number of $`B`$ particles in that site. This corresponds to the average probability for an $`A`$ particle of being involved in the reaction (2) with any of the $`B`$ particles present on the same site. The probabilities $`r_1`$ and $`r_2`$ are related to the reaction rates $`k_1`$ and $`k_2`$ defined in Eqs. (1) and (2). The order parameter of the system is $`\rho _B`$, measuring the density of dynamical entities. For small initial densities $`\rho `$, the system will very likely fall into an absorbing configurations with only frozen $`A`$ particles. For large densities, the system reaches a stationary active state with $`\rho _B0`$.
As we vary $`\rho `$, the system exhibits a continuous transition separating an absorbing phase from an active phase at a critical point $`\rho _c`$. The order parameter is null for $`\rho <\rho _c`$, and follows a power law $`\rho _B(\rho \rho _c)^\beta `$, for $`\rho \rho _c`$. The system correlation length $`\xi `$ and time $`\tau `$, that define the exponential relaxation of space and time correlation functions, both diverge as $`\rho \rho _c`$. In the critical region the system is characterized by a power law behavior, namely $`\xi |\rho \rho _c|^\nu _{}`$ and $`\tau |\rho \rho _c|^\nu _{}`$. The dynamical critical exponent is defined as $`\tau \xi ^z`$, with $`z=\nu _{}/\nu _{}`$. These exponents fully determine the critical behavior of the stationary state of the model (see Ref. ).
We have studied the steady-state properties of the model in $`d=2`$ and $`3`$, by performing numerical simulations for systems with size ranging up to $`L=512`$ and $`L=125`$, respectively. Averages were performed over $`10^410^5`$ independent initial configurations. The values considered for the rates $`r_i`$ are $`r_1=0.1`$ and $`r_2=0.5`$ in $`d=2`$, and $`r_1=0.4`$ and $`r_2=0.5`$ in $`d=3`$. From the finite-size scaling analysis for absorbing phase transitions we obtain the critical point ($`\rho _c=0.3226(1)`$ in $`d=2`$ and $`\rho _c=0.95215(15)`$ in $`d=3`$) and the complete set of critical exponents. A detailed presentation of these results will be reported elsewhere. In Fig. 1 we show as an example the order parameter behavior with respect to the control parameter $`\mathrm{\Delta }=\rho \rho _c`$, from which it is possible to calculate directly the $`\beta `$ exponent. The results obtained in $`d=2`$ and $`3`$ are reported in Tables I and II and compared with the Manna sandpile model in the respective dimension.
In APT it is possible to obtain more information on the critical state by studying the evolution (spread) of activity in systems which start close to an absorbing configuration . In each spreading simulation, a small perturbation is added to an absorbing configuration. It is then possible to measure the spatially integrated activity $`N(t)`$, averaged over all runs, and the survival probability $`P(t)`$ of the activity after $`t`$ time steps. Only at the critical point we have power law behavior for these magnitudes. In the case of many absorbing states, the choice of the initial absorbing state is not unique. There are several methods to perform spreading exponents in this case, and we have followed the technique outlined in Ref.. This procedure amounts to the study of critical spreading with the so-called โnatural initial conditionsโ at $`\rho =\rho _c`$ . The probability distribution $`P_s(s)`$ of having a spreading event involving $`s`$ sites, as well as the the quantities $`N(t)`$ and $`P(t)`$ can thus be measured. At criticality, the only characteristic length is the system size $`L`$, and we can write the scaling forms $`P_s(s)=s^{\tau _s}h(s/L^D)`$, $`N(t)=t^\eta f(t/L^z)`$, and $`P(t)=t^\delta g(t/L^z)`$ . The scaling functions $`f(x)`$, $`g(x)`$ and $`h(x)`$ are decreasing exponentially for $`x1`$, and we have considered that the spreading characteristic time and size are scaling as $`L^z`$ and $`L^D`$, respectively. In this case simulations were performed for systems of size up to $`L=1024`$ in $`d=2`$ and $`L=200`$ in $`d=3`$, averaging over at least $`5\times 10^6`$ spreading experiments. The new scaling exponents $`\tau _s,D,\delta `$, and $`\eta `$ are measured using the now standard moment analysis technique , and a consistency check is executed by performing a data collapse analysis. The resulting exponents are summarized in Tables I and II. As a further consistency check of our results, we have checked that our exponents fulfill all scaling and hyper-scaling relations in standard APT. Despite the apparent diversity in the dynamical rules, we can safely include the SFCRD model and the Manna model in the same universality class. The reported numerical values provide striking evidence for a single universality class, and a further check of the conjecture in Ref..
From a theoretical point of view, the SFCRD allows the microscopic construction of a field theory description that will represent also the critical behavior of all models belonging to the same universality class. The construction of the FT follows the standard steps outlined in the work of Doi, Peliti, and also in Lee and Cardy. This technique consists in recasting the master equation implicit in Eqs. (1) and (2) into a โsecond quantized formโ via a set of creation and annihilation bosonic operators for particles $`A`$ and $`B`$ on each site. At this point it is possible to map the solution of the master equation in a path integral on the density fields weighted by the exponential of a functional action $`S`$. In our case, we can quote the elegant results of Ref., just considering that we have $`D_A=0`$. The action of the FT is thus
$`S`$ $`=`$ $`{\displaystyle }dxdt\{\overline{\psi }[_t+(rD^2)]\psi +\overline{\varphi }[_t\varphi \lambda ^2\psi ]`$ (3)
$`+`$ $`u_1\overline{\psi }\psi (\psi \overline{\psi })+u_2\overline{\psi }\psi (\varphi +\overline{\varphi })`$ (4)
$`+`$ $`v_1\overline{\psi }^2\psi ^2+v_2\overline{\psi }\psi (\psi \overline{\varphi }\overline{\psi }\varphi )+v_3\overline{\psi }\psi \varphi \overline{\varphi }\},`$ (5)
where $`\psi `$ and $`\varphi `$ are auxiliary fields, defined such that their average values coincide with the average density of $`B`$ particles and the total density of particles, respectively, $`\overline{\psi }`$ and $`\overline{\varphi }`$ are response fields, and the coupling constants are related to the reaction rates $`k_i`$. Namely, $`D`$ represents the diffusion coefficient of $`B`$ particles, $`\lambda `$ is initially also proportional to $`D`$, and $`r`$ is the critical parameter that plays the role of the mass in the FT and is related to the difference of the total density with respect to the critical density $`\rho _c`$. The critical point in the mean-field theory is of course at $`r_c=0`$. By standard power-counting analysis, one realizes that the reduced couplings $`u_i/D`$ have critical dimension $`d_c^{(1)}=4`$, while the couplings $`v_i/D`$ have on their part $`d_c^{(2)}=2`$. This means that when applying the renormalization group (RG) and performing a perturbative expansion around the critical dimension $`4`$, one could in principle drop all the couplings $`v_i`$. The critical parameter of this theory is the density of active sites $`\psi `$, while $`\varphi `$ serves just to propagate interactions. We can exploit some symmetry considerations of the FT to relate the physics of the system to the corresponding analytical description. In fact, by neglecting irrelevant terms in the power counting analysis, the action (5) is invariant under the shift transformation
$$\varphi \varphi ^{}=\varphi +\delta ,rr^{}=ru_2\delta ,$$
(6)
where $`\delta `$ is any constant. This symmetry has a very intuitive meaning: If we increase everywhere the density of the system by an amount $`\delta `$, we must be closer to the critical point by an amount proportional to $`\delta `$. In other words, this symmetry represents the conserved nature of the system. It is also interesting to write the set of corresponding Langevin equations (up to irrelevant terms) by integrating out the response fields $`\overline{\psi },\overline{\varphi }`$ in the action $`S`$:
$`_t\psi `$ $`=`$ $`D^2\psi r\psi u_1\psi ^2u_2\psi \varphi +\eta _\psi ,`$ (7)
$`_t\varphi `$ $`=`$ $`\lambda ^2\psi +\eta _\varphi .`$ (8)
Here, $`\eta _\psi `$ and $`\eta _\varphi `$ are noise term with zero mean and correlations $`\eta _\psi (x,t)\eta _\psi (x^{},t^{})=2u_1\psi (x,t)\delta (xx^{})\delta (tt^{})`$, $`\eta _\psi (x,t)\eta _\varphi (x^{},t^{})=u_2\psi (x,t)\delta (xx^{})\delta (tt^{})`$ and $`\eta _\varphi (x,t)\eta _\varphi (x^{},t^{})=0`$. The noise terms have a multiplicative nature , that is the standard form in APT. Note that the $`v_i`$ couplings contribute to the noises correlations with higher order terms. These equations have a very clear physical interpretation. The field $`\varphi `$ is conserved and static, i.e., it only diffuses via the activity of $`B`$ particles, represented by the field $`\psi `$. On its turn, the field $`\psi `$ is locally coupled to the field $`\varphi `$, but is non-conserved. Noticeably, this set of equations recovers the Langevin description that has been proposed on a phenomenological level for stochastic sandpiles, with the extra infromation of the cross-correlation term $`\eta _\psi \eta _\varphi `$ . Indeed, the sandpile model has the same basic symmetries of the present reaction-diffusion model, once the local density field $`\rho `$ is replaced by the local sand-grain (energy) density and the order parameter is identified with the density of toppling sites field. It is then natural to expect that the very same basic structure is reflected in a unique theoretical description. This observation substantiates the existence of a common universality class that embraces stochastic sandpiles, conserved lattice gases and reaction diffusion systems with many absorbing states.
The complete RG analysis of the field theory would allow to extract estimates for the critical exponents to compare with simulations in $`d=2`$ and $`3`$. Unfortunately, several severe technical problems are encountered in this case. In general, as pointed out in Ref., the couplings $`v_i`$ become relevant and should be taken into account in the RG analysis. The importance of the couplings $`v_i`$ can be argued by the change of the energy shift symmetry form, Eq. (6), in the case of the full action, Eq. (5). Second, and more important, is the presence of the singular bare propagator for the field $`\varphi `$, that cannot be regularized by adding a mass term $`m^2\varphi \overline{\varphi }`$, since it will obviously break the symmetry (6). This singular propagator gives rise to divergences in the RG perturbative expansions, and the results of Ref. cannot be extended โtout-courtโ to the limit $`D_A0`$. In particular, some Feynman diagrams in the $`ฯต`$-expansion presented in Refs. are proportional to $`1/D_A`$. Hence, the limit $`D_A0`$ in the theory with $`D_A0`$ is non-analytic; any infinitesimal amount of diffusion in the energy field renormalizes to a finite value, and definitely changes the universality class of the model. Work is in progress to provide a suitable regularization that will allow an $`ฯต`$-expansion calculation of the FT critical exponents.
This work has been supported by the European Network under Contract No. ERBFMRXCT980183. We thank D. Dhar, R. Dickman, P. Grassberger, H. J. Hilhorst, M.A. Muรฑoz, F. van Wijland and S. Zapperi, for helpful comments and discussions.
|
warning/0006/hep-ph0006288.html
|
ar5iv
|
text
|
# References
5 June 2000
Glueballs: A central mystery
Frank E. Close<sup>1</sup><sup>1</sup>1e-mail: F.E.Close@rl.ac.uk
CERN, Geneva, Switzerland
and
Rutherford Appleton Laboratory
Chilton, Didcot, OX11 0QX, England
## Abstract
Glueball candidates and $`q\overline{q}`$ mesons have been found to be produced with different momentum and angular dependences in the central region of $`pp`$ collisions. This talk illustrates this phenomenon and explains the $`\varphi `$ and $`t`$ dependences of mesons with $`J^{PC}=0^{\pm +},1^{++},2^{\pm +}`$. For production of $`0^{++}`$ and $`2^{++}`$ mesons the analysis reveals a systematic behaviour in the data that appears to distinguish between $`q\overline{q}`$ and non-$`q\overline{q}`$ or glueball candidates. An explanation is given for the absence of $`0^+`$ glueball candidates in central production at present energies and the opportunity for their discovery at RHIC is noted.
The idea that glueball production might be favoured in the central region of $`pppMp`$ by the fusion of two Pomerons ( $`IP`$) is over twenty years old . The fact that known $`q\overline{q}`$ states also are seen in this process frustrated initial hopes that such experiments would prove to be a clean glueball source. However, in we noted a kinematic effect whereby known $`q\overline{q}`$ states could be suppressed leaving potential glueball candidates more prominent.
Its essence was that the pattern of resonances produced in the central region of double tagged $`pppMp`$ depends on the vector $`difference`$ of the transverse momentum recoil of the final state protons (even at fixed four momentum transfers). When this quantity ($`dP_T|\stackrel{}{k_{T1}}\stackrel{}{k_{T2}}|`$) is large, ($`O(\mathrm{\Lambda }_{QCD})`$), $`q\overline{q}`$ states are prominent whereas at small $`dP_T`$ all well established $`q\overline{q}`$ are observed to be suppressed while the surviving resonances include the enigmatic $`f_0(1500),f_0(1710)`$ and $`f_0(980)`$.
The data are consistent with the hypothesis that as $`dP_T0`$ all bound states with internal $`L>0`$ (e.g. $`{}_{}{}^{3}P_{0,2}^{}`$ $`q\overline{q}`$) are suppressed while S-waves survive (e.g. $`0^{++}`$ or $`2^{++}`$ glueball made of vector gluons and the $`f_0(980)`$ as any of glueball, or S-wave $`qq\overline{qq}`$ or $`K\overline{K}`$ state). Models are needed to see if such a pattern is natural. Following this discovery there has been an intensive experimental programme in the last two years by the WA102 collaboration at CERN, which has produced a large and detailed set of data on both the $`dP_T`$ and the azimuthal angle, $`\varphi `$, dependence of meson production (where $`\varphi `$ is the angle between the transverse momentum vectors, $`p_T`$, of the two outgoing protons).
The azimuthal dependences as a function of $`J^{PC}`$ and the momentum transferred at the proton vertices, $`t`$, are very striking. As seen in refs. , and later in this paper, the $`\varphi `$ distributions for mesons with $`J^{PC}`$ = $`0^+`$ maximise around $`90^o`$, $`1^{++}`$ at $`180^o`$ and $`2^+`$ at $`0^o`$. Recently, the WA102 collaboration has confirmed that this is not simply a J-dependent effect since $`0^{++}`$ production peaks at $`0^o`$ for some states whereas others are more evenly spread ; $`2^{++}`$ established $`q\overline{q}`$ states peak at $`180^o`$ whereas the $`f_2(1950)`$, whose mass may be consistent with the tensor glueball predicted in lattice QCD, peaks at $`0^o`$ .
In this talk I show how these phenomena arise and in turn expose the extent to which they could be driven, at least in part, by the internal structure of the meson in question and thereby be exploited as a glueball/$`q\overline{q}`$ filter . We find that the $`\varphi `$ dependences of $`0^+`$ and $`1^{++}`$ follow on rather general grounds if a single trajectory dominates the production mechanism. Having thus established the ability to describe the phenomena quantitatively in these cases, we predict the behaviour for $`2^+`$ production and then confront the $`0^{++}`$ and $`2^{++}`$ $`glueball/q\overline{q}`$ sector.
To orient ourselves, think of $`e^+e^{}e^+e^{}M`$ where the essential production dynamics is through $`\gamma \gamma M`$ fusion. The photon can be polarised either $`T`$ ($`\lambda =\pm 1`$) or $`L`$ ($`\lambda =0`$). For $`J^{PC}=0^{++}`$ the resulting structure is $`(Rcos(\varphi ))^2`$ where $`R`$ is equal to the ratio of the longitudinal and transverse production amplitudes for $`\gamma \gamma M`$ and depends on the dynamical structure of the meson, $`M`$. By contrast, parity forbids the production of $`0^+`$ by the fusion of two scalars and also by the longitudinal ($`\mathrm{`}\mathrm{`}L\mathrm{"}`$) components of two vectors. Transverse ($`\mathrm{`}\mathrm{`}T\mathrm{"}`$) components are allowed and so a single amplitude drives the $`\gamma \gamma `$ fusion in production of the $`0^+`$ states. The resulting distribution is predicted to behave like $`sin^2(\varphi )`$.
In ref. it was noted that these distributions are very similar to what is found experimentally in $`pppMp`$ and so a CVC model for the Pomeron was used to confront the data for a range of mesons, $`M`$. The results were very similar, but not identical, to the data, e.g. the $`{}_{}{}^{3}P_{2}^{}`$ $`q\overline{q}`$ states are produced dominantly with $`\lambda =0`$ in $`IP`$ $`IP`$fusion instead of $`\lambda =\pm 2`$ in $`\gamma \gamma `$ fusion. With hindsight the reason is obvious: $`IP`$is not a $`conserved`$ vector current and, in effect, has an intrinsic (and important) scalar component. One effect is that whereas amplitudes for longitudinal $`\gamma `$ emission are suppressed as $`t0`$, the case for the analogous $`IP`$$`grows`$. Suddenly everything began to fit, as summarised in ref. and in the experimental paper โExperimental evidence for a vector-like behaviour of Pomeron exchangeโ . I will now show how the data can be quantitatively described in this simple picture and how characteristic features that may discriminate glueball from $`q\overline{q}`$ states may ensue.
$`J^{PC}=0^+`$
The detailed calculations are described in . Here I shall concentrate on the $`\eta ^{}`$ meson whose production has been found to be consistent with double pomeron exchange . The resulting behaviour of the cross section may be summarised as follows:
$$\frac{d\sigma }{dt_1dt_2d\varphi ^{}}t_1t_2G_{E}^{p}{}_{}{}^{2}(t_1)G_{E}^{p}{}_{}{}^{2}(t_2)\mathrm{sin}^2(\varphi ^{})F^2(t_1,t_2,M^2)$$
where $`\varphi ^{}`$ is the angle between the two $`pp`$ scattering planes in the $`IP`$\- $`IP`$ centre of mass and $`F(t_1,t_2,M^2)`$ is the $`IP`$\- $`IP`$-$`\eta ^{}`$ form factor. We temporarily set this equal to unity; $`pp`$ elastic scattering data and/or a Donnachie Landshoff type form factor can be used as model of the proton- $`IP`$ form factor ($`G_E^p(t)`$).
The WA102 collaboration measures the azimuthal angle ($`\varphi `$) in the $`pp`$ c.m. frame and so we transform the $`\varphi ^{}`$ from the current c.m. frame to $`\varphi `$ for the $`pp`$ c.m. frame. To generalise to real kinematics, we use a Monte Carlo simulation based on Galuga modified for $`pp`$ interactions and incorporating the $`IP`$-proton form factor from ref. .
In order to fit the data we found that the $`IP`$\- $`IP`$-meson form factor $`F(t_1,t_2,M^2)`$ has to differ from unity. If we parametrise $`F^2(t_1,t_2,M^2)`$ as $`exp^{b_T(t_1+t_2)}`$ then we need $`b_T`$ = 0.5 $`GeV^2`$ in order to describe the $`t`$ dependence. Fig. (1a and 1b) compare the final theoretical form for the $`\varphi `$ distribution and the $`t`$ dependence with the data for the $`\eta ^{}`$; (the distributions are well described also for the $`\eta `$ but it has not yet been established that $`IP`$\- $`IP`$ alone dominates the production of this meson).
The $`t_1t_2`$ factors in the cross section arise from the $`TT`$ nature of the amplitude and will be general for the production of any $`0^+`$ meson. Hence for $`0^+`$ states with $`M>>1`$GeV, as expected for the lattice glueball or radial excitations of $`q\overline{q}`$, this dynamical $`t_1t_2`$ factor will suppress the region where kinematics would favour the production. It would be interesting if glueball production dynamics involved a singular $`(t_1t_2)^1`$ that compensated for the transverse $`IP`$factor, as in this case the cross section would stand out. However, we have no reason to expect such a fortunate accident. Hence observation of high mass $`0^+`$ states is expected only to be favourable at extreme energies, such as at RHIC or LHC.
$`J^{PC}=1^{++}`$
In refs. Close and Schuler have predicted that axial mesons are produced polarised, dominantly in helicity one; this is verified by data . The cross section is predicted to have the form
$$\frac{d\sigma }{dt_1dt_2d\varphi ^{}}t_1t_2[\{A(t_1^T,t_2^L)A(t_2^T,t_1^L)\}^2+4A(t_1^T,t_2^L)A(t_1^L,t_2^T)\mathrm{sin}^2(\varphi ^{}/2)]$$
where $`A(t_i,t_j)`$ are the $`IP`$\- $`IP`$-$`f_1`$ form factors. In the models of refs. the longitudinal Pomeron amplitudes carry a factor of $`1/\sqrt{t}`$ arising from the fact that, in the absence of any current conservation for the Pomeron, a longitudinal vector polarisation is not compensated. Thus we make this factor explicit and write $`A(t_i,t_j^L)=\frac{\mu }{\sqrt{t_j}}a(t_i,t_j)`$; hence
$$\frac{d\sigma }{dt_1dt_2d\varphi ^{}}[\{\sqrt{t_1}\sqrt{t_2}\frac{a(t_1^T,t_2^L)}{a(t_1^L,t_2^T)}\}^2+4\sqrt{t_1t_2}\frac{a(t_1^T,t_2^L)}{a(t_1^L,t_2^T)}\mathrm{sin}^2(\varphi ^{}/2)]a^2(t_1^L,t_2^T)$$
In the particular case where the ratio of form factors is unity, this recovers the form used in ref.
$$\frac{d\sigma }{dt_1dt_2d\varphi ^{}}[(\sqrt{t_1}\sqrt{t_2})^2+4\sqrt{t_1t_2}\mathrm{sin}^2(\varphi ^{}/2)]a^2(t_1,t_2)$$
which implies a dominant $`\mathrm{sin}^2(\varphi /2)`$ behaviour that tends to isotropy when suitable cuts on $`t_i`$ are made. This is qualitatively realised (figs. 1e and f of ref. ).
We have parametrised $`a(t_i^T,t_j^L)`$ as an exponential, $`exp^{(b_Tt_i+b_Lt_j))}`$ where $`i,j=1,2`$; $`b_T`$ = 0.5 $`GeV^2`$ was determined from the $`\eta ^{}`$ data above; $`b_L`$ is determined from the overall $`t`$ dependence of the $`1^{++}`$ production and requires $`b_L`$ = 3 $`GeV^2`$. Fig. (2a and b) show the output of the model predictions from the Galuga Monte Carlo superimposed on the $`\varphi `$ and $`t`$ distributions for the $`f_1(1285)`$ from the WA102 experiment.
In addition we have a parameter free prediction of the variation of the $`\varphi `$ distribution as a function of $`|t_1t_2|`$. Fig. (2c and d) show the output of the Galuga Monte Carlo superimposed on the $`\varphi `$ for the $`f_1(1285)`$ for $`|t_1t_2|`$ $``$ 0.2 $`GeV^2`$ and $`|t_1t_2|`$ $``$ 0.4 $`GeV^2`$ respectively. The agreement between the data and our prediction is excellent. Similar conclusions arise for the $`f_1(1420)`$.
$`J^{PC}=2^+`$
The $`J^{PC}`$ = $`2^+`$ states, the $`\eta _2(1645)`$ and $`\eta _2(1870)`$, are predicted to be produced polarised. Helicity 2 is suppressed by Bose symmetry and has been found to be negligible experimentally . The structure of the cross section is then predicted to be
(i) helicity zero: as for the $`0^+`$ case,
$$\frac{d\sigma }{dt_1dt_2d\varphi ^{}}t_1t_2\mathrm{sin}^2(\varphi ^{})$$
(ii) helicity one:
$$\frac{d\sigma }{dt_1dt_2d\varphi ^{}}[\{\sqrt{t_1}\sqrt{t_2}\frac{a(t_1^T,t_2^L)}{a(t_1^L,t_2^T)}\}^2+4\sqrt{t_1t_2}\frac{a(t_1^T,t_2^L)}{a(t_1^L,t_2^T)}\mathrm{cos}^2(\varphi ^{}/2)]a^2(t_1^L,t_2^T)$$
which is as the $`1^{++}`$ case except for the important and significant change from $`\mathrm{sin}^2(\varphi ^{}/2)`$ to $`\mathrm{cos}^2(\varphi ^{}/2)`$.
The uncompensated factor of $`t_1t_2`$ in the helicity zero component tends to suppress this kinematically under the conditions of the WA102 experiment. Indeed, WA102 find that helicity one alone is able to describe their data ; this is in interesting contrast to $`\gamma \gamma \eta _2(Q\overline{Q})`$ in the non-relativistic quark model where the helicity-one amplitude would be predicted to vanish . We shall concentrate on this helicity-one amplitude henceforth.
The results of the WA102 collaboration for the $`\eta _2(1645)`$ are shown in fig. (3a and b). The distribution peaks as $`\varphi 0`$, in marked contrast to the suppression in the $`1^{++}`$ case (fig. 2a).
Integrating our formula over $`\varphi `$, with the same approximations as previously, implies
$$\frac{d\sigma }{dt_1dt_2}(t_1+t_2)(exp^{(b(t_1+t_2)})$$
and, in turn, that
$$\frac{d\sigma }{dt}(1+bt)(exp^{bt})$$
(1)
This simple form compares remarkably well with WA102 who fit to $`\alpha e^{b_1t}+\beta te^{b_2t}`$; our prediction (eq. 1) implies that $`b_1b_2`$ and that $`\beta /\alpha b`$ and WA102 find for the $`\eta _2(1645)`$ $`b_1=6.4\pm 2.0;b_2=7.3\pm 1.3`$ and $`\beta =2.6\pm 0.9`$, $`\alpha =0.4\pm 0.1`$
Performing the detailed comparison of model and data via Galuga, as in the previous examples, leads to the results shown in fig. (3a and b) for the $`\eta _2(1645)`$; the $`\eta _2(1870)`$ results are qualitatively similar. Bearing in mind that there are no free parameters, the agreement is remarkable. Indeed, the successful description of the $`0^+`$, $`1^{++}`$ and $`2^+`$ sectors, both qualitatively and in detail, sets the scene for our analysis of the $`0^{++}`$ and $`2^{++}`$ sectors where glueballs are predicted to be present together with established $`q\overline{q}`$ states. Any differences between data and this model may then be a signal for hadron structure, and potentially a filter for glue degrees of freedom.
Before turning to the $`0^{++},2^{++}`$ channels with glueball interest, it is worth summarising exactly what we have assumed and what we have described, parameter free.
For the production of $`J^{PC}`$ = $`0^+`$ mesons we have predicted the $`\varphi `$ dependence and the vanishing cross section as $`t0`$ absolutely and have fitted the $`t`$ slope in terms of one parameter, $`b_T`$. For the $`J^{PC}`$ = $`1^{++}`$ mesons we predict the general form for the $`\varphi `$ distribution: it is in this channel for the first time that the non-conserved nature of the $`IP`$first manifests itself. The polarisation of the $`1^{++}`$ is also natural. By fitting the $`t`$ slope we obtain the parameter $`b_L`$; this then gives a parameter free prediction for the variation of the $`\varphi `$ distribution as a function of $`t`$ which agrees with the data. With parameters now fixed, we obtain absolute predictions for both the $`t`$ and $`\varphi `$ dependences of the $`J^{PC}`$ = $`2^+`$ mesons which are again in accord with the data when helicity 1 dominance is imposed.
The message is that the production of the unnatural spin-parity states, $`0^+,1^{++},2^+`$, is driven by the non-conserved vector nature of the exchanged $`IP`$; it is not immediately affected by the internal structure of the produced meson. In particular, it is not sensitive to whether the mesons are glueballs or $`q\overline{q}`$.
Now I shall look at the $`0^{++}`$ and $`2^{++}`$ sector where glueballs are expected as well as $`q\overline{q}`$. Here we shall find that the production topologies do depend on the internal dynamics of the produced meson and as such may enable a distinction between $`q\overline{q}`$ and exotic, glueball, states.
$`J^{PC}=0^{++}`$ and $`2^{++}`$
In contrast to the $`0^+`$ case, where parity forbade the LL amplitude, in the $`0^{++}`$ case both $`TT`$ and $`LL`$ can occur. Hence there are two independent form factors $`A_{TT}(t_1,t_2,M^2)`$ and $`A_{LL}(t_1,t_2,M^2)`$. For $`0^{++}`$ and the helicity zero amplitude of $`2^{++}`$ (which experimentally is found to dominate ) the angular dependence of scalar meson production will be
$$\frac{d\sigma }{dt_1dt_2d\varphi ^{}}G_{E}^{p}{}_{}{}^{2}(t_1)G_{E}^{p}{}_{}{}^{2}(t_2)[1+\frac{\sqrt{t_1t_2}}{\mu ^2}\frac{a_T}{a_L}e^{(b_Lb_T)(t_1+t_2)/2}\mathrm{cos}(\varphi ^{})]^2e^{b_L(t_1+t_2)}$$
(2)
where we have written $`a_L(t)=a_Le^{(b_Lt/2)}`$ and $`a_T(t)=a_Te^{(b_Tt/2)}`$ with $`b_{L,T}`$ fixed to the values found earlier. The ratio $`a_T/a_L`$ can be positive or negative, or in general even complex.
Eq.(2) predicts that there should be significant changes in the $`\varphi `$ distributions as $`t`$ varies. When $`\frac{\sqrt{t_1t_2}}{\mu ^2}a_T/a_L\pm 1`$, the $`\varphi `$ distribution will be $`\mathrm{cos}^4(\frac{\varphi }{2})`$ or $`\mathrm{sin}^4(\frac{\varphi }{2})`$ depending on the sign. Indeed data on the enigmatic scalars $`f_0(980)`$ and $`f_0(1500)`$ show a $`\mathrm{cos}^4(\frac{\varphi }{2})`$ behaviour when $`\sqrt{t_1t_2}0.1`$ GeV<sup>2</sup>, changing to $`\mathrm{cos}^2(\varphi )`$ when $`\sqrt{t_1t_2}0.3`$ GeV<sup>2</sup> .
The overall $`\varphi `$ dependences for the $`f_0(1370)`$, $`f_0(1500)`$, $`f_2(1270)`$ and $`f_2(1950)`$ can be described by varying the quantity $`\mu ^2a_L/a_T`$. Results are shown in fig. 4. It is clear that these $`\varphi `$ dependences discriminate two classes of meson in the $`0^{++}`$ sector and also in the $`2^{++}`$. The $`f_0(1370)`$ can be described using $`\mu ^2a_L/a_T`$ = -0.5 $`GeV^2`$, for the $`f_0(1500)`$ it is +0.7 $`GeV^2`$, for the $`f_2(1270)`$ it is -0.4 $`GeV^2`$ and for the $`f_2(1950)`$ it is +0.7 $`GeV^2`$.
It is interesting to note that we can fit these $`\varphi `$ distributions with one parameter and it is primarily the sign of this quantity that drives the $`\varphi `$ dependences. Understanding the dynamical origin of this sign is now a central issue in the quest to distinguish $`q\overline{q}`$ states from glueball or other exotic states.
Acknowledgements
This is based on work performed in collaborations with A.Kirk and G.Schuler and is supported, in part, by the EU Fourth Framework Programme contract Eurodafne, FMRX-CT98-0169.
Figures
Figure 1
Figure 2
Figure 3
Figure 4
|
warning/0006/hep-th0006127.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
####
The general relativity is invariant under general coordinate transformations like $`xx^{}=x+\xi (x)`$. One can still define a convenient covariant derivative under a general transformation of coordinate as $`๐_\mu \xi _\nu =_\mu \xi _\nu +\mathrm{\Gamma }_{\mu \nu }^\lambda \xi _\lambda `$ where $`\xi _\mu (x)`$ is the infinitesimal parameter and $`\mathrm{\Gamma }_{\mu \nu }^\lambda `$ is the Christoffel symbol or โconnectionโ.
The infinitesimal variation of metric under the coordinate transformation can be written as $`\delta g_{\mu \nu }(x)=๐_\mu \xi _\mu (x)๐_\nu \xi _\mu (x)`$.
Itโs easy to verify that for weak field approximation, $`g_{\mu \nu }(x)=\eta _{\mu \nu }+kh_{\mu \nu }`$ with a redefinition of the $`\xi _\mu (x)`$ parameter. We obtain the gauge transformation imposed on the field $`h_{\mu \nu }(x)`$ such as infinitesimal version in the Minkowsky space under GCTโs (general transformation of coordinates).
How do we consider the interaction between gravitation and other fields? The general recipe is to introduce new fields and to define an appropriate covariant derivative that includes such an interaction.
In general there are two ways to follow: Firstly, we can use the coupling between gravitation and other fields. In this case the gravitational field is only a background for other fields. The scenario is geometrical as for gravity which acts as a background field. There is no โinteractionโ as is the case for the electromagnetic theory, quantum chromodynamics or even in the perturbative quantum gravity. On the other hand we can introduce a flat space time and attack the problem using the perturbative approach. If we wish to interpret the fields as particles associated with that field we need to introduce a local Lorentz manifold to bring in a specific interactions. It is true, in particular if we wish to analyse the interaction between gravitation and spinor fields. The interaction cannot be introduced directly in curved space time for the spinor field. Itโs necessary to look for a local Lorentz manifold rather than the global manifold as in space time.
This point is emphasized, in particular, for the special case of Liouville-Vlasov equation. We believe that the correct way to introduce the effect of gravitational field in that equation is by defining an appropriate covariant derivative in a local Lorentz manifold against the point of view that one can define the gravitational effect on Vlasov equation directly in curved space time. The latter is written as follows.
We analyse the interaction between gravitation and spinor fields using the vierbein field and the local Lorentz group as our support. As a second example we present the coupling between matter-Maxwell and gravitation fields. The following example is the $`U(1)`$ gauge group generalized to the case of, $`SO(N)`$ model coupled to gravity and spinor-vector field (Rarita-Schwinger). In all cases the interaction is seen on a local Lorentz manifold where it is possible to find a spin connection, and in some cases, to find the โtorsionโ associated with such a model.
Finally, the Liouville-Vlasov equation is discussed and the effect of gravitational field on that equation is analysed. We argue here that it is impossible to define a covariant derivative for the Vlasov equation if $`\psi (q,p)`$ means the Wigner function associated with the spinor field. The fundamental reason for it is that the symmetry group of general relativity is incompatible with the presence of spinors ; but we can do that only in a local Lorentz manifold where both the Vierbein and spin connection are defined. We start by remembering that the concept of spin $`1/2`$ field makes sense only in a tangent flat space . One can write the action for interaction between Diracโs field and gravitation as
$$S=d^4xe\left[e^{\mu a}\overline{\psi }i\gamma _aD_\mu \psi m\overline{\psi }\psi \right]$$
(1)
where
$$\gamma ^\mu =e_a^\mu (x)\gamma ^a$$
(2)
and
$$D_\mu \psi _\alpha =_\mu \psi _\alpha +\frac{i}{8}B_\mu ^{ab}[\gamma _a\gamma _b]_{\alpha \beta }\psi _\beta $$
(3)
with $`e_a^\mu `$ are the Vierbein field with two indices; one space time index a, $`\mu `$, varying from zero to three and a local Lorentz group index, a, of a local manifold. The $`\gamma ^\mu `$ are the Diracโs matrices, $`\gamma ^a`$ meaning the local Dirac matrices and $`B_\mu ^{ab}`$ being the spin connection.
The interaction between gravitation and fermionic field can be seen immediately from eq. (3) because of the appropriate definition of a covariant derivative $`D_\mu `$ for a local Lorentz group.
The metric can be written as
$$g_{\mu \nu }=e_\mu ^ae_\nu ^b\eta _{ab}.$$
(4)
Thus, the โeโ in eq. (1) means the determinant of the metric. The Dirac equation coupled to gravitation is given by
$$(e^{\mu a}i\gamma _aD_\mu m)\psi (x)=0$$
(5)
The spin connection in (3) can be completly fixed by the coefficient of non holomicity as
$$B_\mu ^{ab}=\frac{1}{2}e_\mu ^c(\mathrm{\Omega }_{cda}+\mathrm{\Omega }_{acd}\mathrm{\Omega }_{dac})$$
(6)
where
$$\mathrm{\Omega }_{cda}=e_c^\rho e_j^\sigma (_\rho e_{\sigma a}_\sigma e_{\rho a})$$
(7)
The equation for $`e_\mu ^a`$ field, $`R_{\mu a}\frac{1}{2}e_{\mu a}R=0`$ fixes $`e_\mu ^a(x)`$ exactly; then we have that $`B_\mu ^{ab}`$ shall be fixed too.
The gravitational degree of freedom is carried by Vierbein. All information about spins from gravitational field will be given by $`e_\mu ^a(x)`$.
One may see clearly that the kinetic term of the fermionic lagrangean contributes to the equation of motion for the spin connection and we can show that the presence of the fermionic field generates torsion given as
$$T_{\mu \nu }^aD_\mu (ee_{[a}^\mu e_{b]}^\nu )\overline{\psi }\gamma [\gamma ,\gamma ]\psi .$$
(8)
where $`[a,b]`$ here means symmetrization and the right hand side represents the fermionic density.
## The interaction between matter-Maxwell and gravitational fields
####
The action for interaction between matter-Maxwell and gravitational fields is given as
$$S_{MatGrav}^{Maxwell}=d^4xe[e^{\mu a}\overline{\psi }i\gamma _a_\mu \psi m\overline{\psi }\psi ]$$
(9)
where $`_\mu `$ is a covariant derivative of gravitation and the gauge is given by
$$_\mu \psi =D_\mu +igqA_\mu $$
(10)
with $`D_\mu `$ being the covariant derivative as in eq. (3), $`A_\mu (x)`$ is the vector potential, $`q`$ is a coupling constant associated with the $`U(1)`$ symmetry and $`g`$ is another coupling constant linking the local Lorentz group. The action is invariant under local Lorentz group, GCTโs (transformation coordinates group) and $`U(1)`$ symmetry simultaneously.
A term is needed that gives the dynamics for the gauge degree of freedom (Maxwell term); so the complete action which couples Maxwell-Dirac-gravitation is written as
$$S_{DiracGrav}^{Maxwell}=d^4xe[\frac{1}{4}F_{\mu \nu }F^{\mu \nu }+e^{\mu a}\overline{\psi }i\gamma _a_\mu \psi m\overline{\psi }\psi ]$$
(11)
## The $`U(1)`$ case generalized
####
Consider the scalar and spinor field matter fields coupled to gravitation and Maxwell field. The lagrangean is given as
$$=(_\mu \phi ^{}^\mu \phi m^2\phi ^{}\phi )+\overline{\psi }(i\gamma ^\mu _\mu M)\psi +\frac{\lambda }{4}(\phi ^{}\phi )^2.$$
(12)
To obtain a covariant lagrangean the convenient covariant derivative is introduced as $`_\mu _\mu `$, where
$$_\mu \psi =\left(_\mu +\frac{i}{8}B_\mu ^{ab}[\gamma _a,\gamma _b]+igQA_\mu (x)\right)\psi $$
(13)
Thus, the complete action for Maxwell, matter and gravitational field is
$$S_{matgrav}^{Maxwell}=d^4xe\left[(_\mu \phi ^{})(^\mu \phi )V(\phi ^{},\phi )\right]+ee^{\mu a}\overline{\psi }(i\gamma _a_\mu \psi M)\psi $$
(14)
where $`V(\phi ^{}\phi )=m^2\phi ^{}\phi `$ is the potential term and $`Q`$ means the generalized charge.
The equation of motion is immediately obtained as
$$^\mu _\mu \phi m^2\phi +\mathrm{}=0$$
(15)
and the dynamics is verified on local Lorentz manifold again.
## Model $`SO(N)`$ coupled to gravity
####
The interest in this model is formal. There appear magnetic monopoles of tโHooft-Polyakov coupling to gravitational field.
Suppose that $`\phi _a`$ are scalars fields in the representation of $`SO(N)`$ and $`a=1,2,3\mathrm{}N`$.
Thus,
$$\phi _a{}_{}{}^{}=R_{ab}\phi _b$$
(16)
where $`R_{ab}`$ is the transformation matrix and it satisfies the condition $`R^+R=1`$.
The lagrangean satisfying the invariance under global $`SO(N)`$ is given by
$$=\frac{1}{2}_\mu \phi _a^\mu \phi _a\frac{1}{2}m^2\phi _a\phi _a.$$
(17)
the $`SO(N)`$ will be calibrated by equation (16) and we need to change again the derivative $`_\mu `$ to a new covariant derivative $`_\mu `$ written as
$$_\mu \phi _a=(_\mu \delta _{ab}+igA_\mu ^I(G_I)_{ab})\phi _b$$
(18)
where $`G_I`$ are generators of $`SO(N)`$ group and $`A_\mu ^I(x)`$ are the non abellian vector potentials. The number of generators being written as $`I=1\mathrm{}{\displaystyle \frac{N(N1)}{2}}`$.
The complete action for this case can be written as
$`S_{matgrav}^{Y.M}={\displaystyle }d^4xe[{\displaystyle \frac{1}{4}}F_{\mu \nu I}F^{\mu \nu I}+{\displaystyle \frac{1}{2}}_\mu \phi _a^\mu \phi _a{\displaystyle \frac{1}{2}}m^2\phi _a\phi _a+`$
$`{\displaystyle \frac{\lambda }{4!}}(\phi _a\phi _b)^2+fR\phi _a\phi _a]`$ (19)
where $`f`$ means the dimensionless coupling constant, $`R`$ in the last term is the scalar curvature and $`F_{\mu \nu I}`$ represents the field strength given by
$$F_{\mu \nu I}=_\mu A_{\nu I}_\nu A_{\mu I}+gf_{IJK}A_{\mu J}A_{\nu K}$$
(20)
## Spinor-vector fields
####
We define now $`\psi _{a\alpha }`$ as a spinor-vector field with $`\alpha =1,2,3,4`$ and $`a=0,1,2,3`$. The transformation law for Rarita-Schwinger field can be written as
$$\delta \psi _{a\alpha }=w_a^b\psi _{b\alpha }+\frac{1}{2}w^{mn}[\gamma _m,\gamma _n]_{\alpha \beta }\psi _{\alpha \beta }$$
(21)
where $`w_a^b`$ are continuously varying parameters. The first part on right hand side being responsible for spin-1 and the second part being associated with spin field $`1/2`$ and $`3/2`$.
The spin $`1/2`$ can be eliminated consistently by a symmetry of the free lagrangean . Thus, $`\psi _{a\alpha }`$ shall describe a pure spin $`3/2`$.
It appears in supergravity as a fermionic mediator of the gravitational interaction.
The lagrangean which describes a free spinor-vector field can be given as
$$_{R.S.}=\frac{1}{2}\epsilon ^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\gamma _\nu _\rho \psi _\sigma $$
(22)
with $`\epsilon ^{\mu \nu \rho \sigma }`$ being the Levi-Civita tensor and $`\gamma _5=\gamma _0\gamma _1\gamma _2\gamma _3`$ is a Diracโs matrix. We can write eq. (22) in the component form as
$$_{R.S.}=\frac{1}{2}\epsilon ^{\mu \nu \rho \sigma }\overline{\psi }_{\mu \alpha }(\gamma _5\gamma _\nu )_{\alpha \beta }_\rho \psi _{\sigma \beta }$$
(23)
The local invariance $`U(1)`$ that permits the elimination of spin $`1/2`$ components is given as
$$\delta \psi _{\mu \alpha }=_\mu \chi $$
(24)
where $`\chi `$ is a spinor field. As in the case of the fermionic field the Rarita-Schwinger field generates torsion .
The interaction between gravitation and the spinor-vector field is shown following the same recipe as in earlier cases so that
$$_{R.S.}^{gravity}=\frac{e}{2}\epsilon ^{\mu \nu \rho \sigma }\overline{\psi }_{\mu \alpha }(\gamma ^5\gamma _\nu )_{\alpha \beta }D_\rho \psi _{\sigma \beta }$$
(25)
where
$$D_\rho \psi _{\sigma \beta }=_\rho \psi _{\sigma \beta }+\frac{i}{8}B_\rho ^{ab}[\gamma _a,\gamma _b]_{\beta \gamma }\psi _{\sigma \gamma }$$
(26)
It is clear that the only difference between eq. (25) and the other cases is that here $`\psi `$ has two indices. One can show again that the spinor-vector field generates torsion again as $`T_{\mu \nu }^a\overline{\psi }\gamma _5\gamma [\gamma ,\gamma ]\psi `$ forming a spinor condensity exactly the same way as given in (8).
## The Liouville-Vlasov equation with the gravitational interaction
####
Finally, we obtain the Liouville-Vlasov equation for the Wigner function of a โspinor fieldโ coupled to a gauge field with the field strength tensor $`F^{\mu \nu }`$ as
$$(p_\mu F^{\mu \nu }D_\nu +P^\mu D_\mu )\psi (q,p)=0$$
(27)
where $`\psi (q,p)`$ is the Wigner function. Here $`D_\mu `$ cannot be a covariant derivative including the curved space-time spin connection unlike the case discussed in .
The reason for this is that the coordinate tranformation group of general relativity is incompatible with the presence of spinor field and so, incompatible with the appearance of spin connection $`B_\mu ^{ab}`$ in curved space-time .
We can introduce an appropriate covariant derivative in eq. (27) the same way as in preceding cases. So, considering the tangent space or local Lorentz group we do the same use of eq. (3) and in all analyzed cases, obtained for $`D_\mu `$ in eq. (27) a covariant derivative like
$$D_\mu \psi =_\mu \psi +\frac{i}{8}B_\mu ^{ab}[\gamma _a,\gamma _b]\psi $$
(28)
where $`\psi `$ is the Wigner function associated with a spinor field coupled to a gauge field in a local Lorentz manifold (tangent space). It is well-known that $`\psi ^{}(q,p)\psi (q,p)`$ can be interpreted as the classical distribution function in the relativistic phase space.
We observe that if we wish to include the real effects from gravitational field in an approach of general relativity together with the kinetic theory some troubles will appear. For example: it is clear โthat the treatment of transport theory in a curved space-time background is hidered by the fact that a definition of the Wigner function $`\psi (q,p)`$ depends on the use of the Fourier transform of a space-time correlation function with a translated argumentโ.
Fourier transformation (nor translation) are globally available in general curved spacetime. On the other hand the symmetry group of general relativity is not compatible with the presence of spinors or spin connection field if the Wigner function of a spinor field in a curved space time has to be found.
However, as discussed in , the construction of Wigner function $`\psi (q,p)`$ is still meaningful if carried out in the tangent space as in local Lorentz manifold as applied in our case. The problem is that the correlation between two points on that manifold is by an exponential map . Only this way, we can introduce the fermionic field interacting with the gravitation and to define an appropriate covariant derivative containing the spin connection.
### Acknowledgements:
####
I would like to thank the Department of Physics, University of Alberta for their hospitality. This work was supported by CNPq (Governamental Brazilian Agencie for Research.
I would like to thank also Dr. Don N. Page for his kindness and attention with me at Univertsity of Alberta and Dr. Ademir Santana from Federal University of Bahia by fruitful discussions.
|
warning/0006/cond-mat0006393.html
|
ar5iv
|
text
|
# Hofstadter butterfly and integer quantum Hall effect in three dimensions
## Abstract
For a three-dimensional lattice in magnetic fields we have shown that the hopping along the third direction, which normally tends to smear out the Landau quantization gaps, can rather give rise to a fractal energy spectram akin to Hofstadterโs butterfly when a criterion, found here by mapping the problem to two dimensions, is fulfilled by anisotropic (quasi-one-dimensional) systems. In 3D the angle of the magnetic field plays the role of the field intensity in 2D, so that the butterfly can occur in much smaller fields. The mapping also enables us to calculate the Hall conductivity, in terms of the topological invariant in the Kohmoto-Halperin-Wuโs formula, where each of $`\sigma _{xy},\sigma _{zx}`$ is found to be quantized.
Among the effects of magnetic fields on electronic states, one of the most bizzare is Hofstadterโs butterfly. Namely, when a two-dimensional (2D) periodic system is put into a magnetic field, the gap not only appears between the Landau levels, but a series of gaps appear in a selfsimilar fashion as shown by Hofstadter. The butterfly refers to an energy spectrum against the magnetic flux, $`\varphi `$, penetrating a unit cell in units of the flux quantum $`\varphi _0=h/e`$. Usually the butterfly is conceived to be a phenomenon specific to 2D.
Here we raise a question: can we have something like Hofstadterโs butterfly in spite of, or even because of, a three-dimensionality (3D)? This may at first seem quite unlikely, since the usual derivation of the butterfly relies on the two-dimensionality of the system, so that a motion along the third direction ($`z`$) should tend to wash out the butterfly gaps as well as Landau level gaps. Several authors have extended Hofstadterโs problem to 3D in the last decade , and subbands are indeed shown to overlap or touch with each other. Here we systematically look for the possibility of butterflies (i.e., recursive and fractal gaps) in 3D.
If we do have a butterfly, then we can proceed to question how the integer quantum Hall effect should look like on the butterfly. If one examines a theoretical reasoning from which the quantization in the Hall conductivity is deduced in the usual quantum Hall system, the essential ingredient is the presence of a gap in the energy spectrum. This was indicated in a gauge argument by Laughlin, and elaborated by Thouless, Kohmoto, Nightingale, and den Njis when a periodic potential exists. There the quantized Hall conductivity when the Fermi energy, $`E_F`$, lies in a butterfly gap is identified to be a topological invariant characterizing the position of gaps.
For 3D Kohmoto, Halperin, and Wu have shown, following the line of the 2D work, that if there is an energy gap in a 3D system, then an integer quantum Hall effect should result as long as $`E_F`$ lies within a gap. Montambaux and Kohmoto have actually calculated the Hall conductivity in a case where a third-direction hopping opens some gaps. If butterflies are realized in 3D, we can move on to the systematics of the quantum Hall numbers.
In the present paper we point out that, first, an analog of Hofstadterโs butterfly does indeed exist specific to 3D under certain criterion that is fulfilled by anisotropic (quasi-1D) tight-binding lattices in 3D. In this case the butterfly plot refers to energy versus angle of the magnetic field. This is obtained by mapping the 3D system to a 2D system. Remarkablly the mapping dictates that the ratio of the magnetic fluxes penetrating two facets of the unit cell plays the role of the magnetic flux in 2D, so that the field intensity $`B`$ does not have to be strong to realize the butterfly.
More importantly the mapping enables us to systematically calculate the Hall conductivity for the 3D butterfly via identifying the topological invariant in the Kohmoto-Halperin-Wuโs formula. We have found that each of $`\sigma _{xy},\sigma _{zx}`$ is quantized in 3D.
Our model is non-interacting tight-binding electrons in a uniform magnetic field $`๐ฉ`$ described by the Hamiltonian,
$$=\underset{i,j}{}(t_{ij}e^{i\theta _{ij}}c_i^{}c_j+\mathrm{h}.\mathrm{c}.),$$
(1)
in standard notations, where the summation is taken over nearest-neighbor sites with $`t_{ij}=t_x,t_y,t_z`$ along $`x,y,z`$, respectively, $`\theta _{ij}=(e/\mathrm{})_i^j๐จ๐๐`$ is the Peierls phase. We first recapitulate the 2D case, because a key in this work is a correspondence between 2D and 3D. In 2D with the Landau gauge $`๐จ=(0,Bx)`$, $`y`$ is a cyclic coordinate, so that the wave function becomes $`\psi _{lm}=e^{i\nu _ym}F_l`$, where $`(l,m)`$ labels $`(x,y)`$ coordinates, and $`\nu _y`$ is the Bloch wave number along $`y`$. The Schrรถdinger equation then takes a form of Harperโs equation,
$$t_x(F_{l1}+F_{l+1})\mathrm{\hspace{0.17em}2}t_y\mathrm{cos}(2\pi \varphi l+\nu _y)F_l=EF_l,$$
(2)
where $`\varphi =Bab/\varphi _0`$ is the number of flux quanta penetrating a unit cell $`=a\times b`$. While the energy spectrum becomes a butterfly for the ordinary isotropic case, $`t_x=t_y`$, the gaps are rapidly smeared out as the anisotropy is introduced, $`t_y/t_x0`$, since the potential term (the cosine function above) weakens. Since $`t_x`$ and $`t_y`$ appear on an equal footing, the condition for the appearance of the butterfly in 2D is $`t_xt_y`$.
Harperโs equation in 3D can be derived in a similar way. For simplicity we consider a simple cubic lattice in a magnetic field $`๐ฉ=(0,B\mathrm{sin}\theta ,B\mathrm{cos}\theta )`$ assumed to lie in the $`yz`$ plane. The vector potential is then $`๐จ=(0,Bx\mathrm{cos}\theta ,Bx\mathrm{sin}\theta )`$, so that $`y,z`$ are cyclic and the wave function becomes $`\psi _{lmn}=e^{i\nu _ym+i\nu _zn}F_l`$, where $`(l,m,n)`$ labels $`(x,y,z)`$. The Schrรถdinger equation is
$`t_x(F_{l1}+F_{l+1})[2t_y\mathrm{cos}(2\pi \varphi _zl+\nu _y)`$ (3)
$`+2t_z\mathrm{cos}(2\pi \varphi _yl+\nu _z)]F_l=EF_l,`$ (4)
where two periodic potentials are now superposed. Here $`\varphi _y(\varphi _z)`$ is the number of flux quanta penetrating the side of a unit cell ($`=a\times b\times c`$) normal to $`y(z)`$ (inset of Fig.3(a)).
Although the spectrum in 3D does not in general have many gaps (aside from the trivial Bragg-reflection gaps), we first notice that a butterfly-like structure does emerge for certain choices of $`(t_x,t_y,t_z)`$, as typically displayed in 1(a) for $`(t_x,t_y,t_z)=(1,0.1,0.1)`$, a quasi-1D system. The spectrum is plotted against the angle $`\theta `$ of a magnetic field $`(\varphi _y,\varphi _z)=0.2(\mathrm{sin}\theta ,\mathrm{cos}\theta )`$ with $`b=c`$ assumed here. A structure akin to the 2D butterfly is seen in the bottom (or at the top) of the whole spectrum. One might consider this as a 2D butterfly surviving the third-direction hopping, but this is wrong as is evident from Fig.1(b), where we turn off $`t_z`$ to find that the spectrum coalesces to a series of broadened Landau levels. So we are talking about the butterfly specific to 3D rather than a remnant of a 2D butterfly.
We first explore the mechanism why the butterfly appears in 3D. The doubly periodic potential in the 3D Harper equation comprises $`V^{(1)}(l)t_y\mathrm{cos}(2\pi \varphi _zl+\nu _y),V^{(2)}(l)t_z\mathrm{cos}(2\pi \varphi _yl+\nu _z)`$. We assume that their periods, $`1/\varphi _z`$ and $`1/\varphi _y`$, are much greater than the lattice constant ($`\varphi _z,\varphi _y1`$). We also assume that
$$t_y\varphi _zt_z\varphi _y,$$
(5)
which amounts to an assumption that the local peaks and dips of the total potential $`V^{(1)}+V^{(2)}`$ is primarily that of $`V^{(1)}`$.
One can then regard the potential minima of $`V^{(1)}`$ as โsitesโ, which we call โwellsโ to distinguish from the original sites. The wells are separated by $`1/\varphi _z`$ and feel the slowly-varying $`V^{(2)}`$. Since each well contains many original sites due to the first assumption, we can talk about bound states for the well in the effective-mass approach. If wells are deep enough, several bound states appear and each state forms a tight-binding band (i.e., Landau band), and the equation (4) reduces to
$`t^{}(J_{l^{}1}+J_{l^{}+1})\mathrm{\hspace{0.17em}2}t_z\mathrm{cos}[2\pi (\varphi _y/\varphi _z)l^{}`$ (6)
$`+(\varphi _y/\varphi _z)\nu _y+\nu _z]J_l^{}=EJ_l^{}.`$ (7)
Here $`t^{}`$ is a transfer integral between neighboring wells labelled by $`l^{}`$, $`J_l^{}`$ the โeffective-massโ wave function, and the cosine term is the value of $`V^{(2)}`$ at each minimum of $`V^{(1)}`$. The reduced equation has exactly the same form as that of the 2D system, eqn.(2) if we translate
$$3\mathrm{D}:(t_x,t_y,t_z,\varphi _y,\varphi _z)2\mathrm{D}:(t^{},t_z,\varphi _y/\varphi _z).$$
(8)
Since the butterfly is a hallmark of an isotropic 2D case, one can predict that the subband in 3D for which $`t^{}t_z`$ should exhibit a butterfly.
We can estimate $`t^{}`$ by applying the effective-mass approximation to Harperโs equation (4). We first convert the equation (when there is $`V^{(1)}`$ alone) to a differential equation for a continuous variable $`\stackrel{~}{l}2\pi \varphi _zl`$, which turns out to contain a combination $`t_y/\varphi _z^2`$ only (with $`t_x=1`$, a unit of energy). Since $`t^{}`$ is a matrix element of $`V^{(1)}t_y`$, we have a simple scaling law,
$$t^{}=2t_yf\left(\frac{t_y}{\varphi _z^2}\right).$$
(9)
Note that the value of $`t^{}`$ differs from one tight-binding band to another, where middle bands, with weaker binding, have larger $`t^{}`$.
We have then numerically calculated $`t^{}(t_y,\varphi _z)`$ for the lowest band. With the scaling, the $`t_y`$-dependence of $`t^{}`$ for a particular $`\varphi _z`$ provides the whole dependence, shown in Fig.2. If we plug in the condition for the butterfly, $`t_zt^{}`$, the plot indicates how to adjust $`\varphi _z`$ to have a butterfly for given $`(t_y,t_z)`$. One can immediately find that the butterfly is restricted to the case with $`t_y,t_z1(=t_x)`$, i.e., quasi-1D systems. This is because $`\varphi _z`$ becomes too large to satisfy $`\varphi _z1`$ in the most region of $`t_y1`$ or in the region $`t_z1`$ (out of the plot).
This plot shows that the above example, $`(t_x,t_y,t_z)=(1,0.1,0.1)`$, is indeed a right choice, for which we can show that $`t^{}=0.05t_z(=0.1)`$ for the lowest subband. Incidentally, the butterfly is symmetric in this example, which is an accident for $`t_y=t_z`$: in Harperโs equation $`V^{(1)}`$ and $`V^{(2)}`$ exchange roles at $`\theta =45^{}`$ for $`t_y=t_z`$. Around $`45^{}`$, or more generally around $`t_y\varphi _zt_z\varphi _y`$, the above argument breaks down but a clear structure remains. We can explain this as follows. When $`t_y\varphi _zt_z\varphi _y`$, $`V^{(1)}+V^{(2)}`$ exhibits a beat so that the barrier height separating the wells varies from place to place, which implies that $`t^{}`$ varies from place to place. However, a change in the wall height ($`t_y`$) changes $`t^{}`$ only slightly, since $`t^{}`$ has a broad peak in Fig.2.
Thus in 3D the effective flux in eqn.(8) is a ratio $`\varphi _y/\varphi _z`$. In 2D by contrast, a butterfly requires $`\varphi O(1)`$, i.e., $`B`$ has to be impossibly large ($`10^5`$T for $`a=2`$ร
). Would this render the 3D butterfly experimentally feasible? In principle, Fig.2 shows that there exists appropriate $`(t_y,t_z)`$ no matter how small $`\varphi _z`$ may be. In practice, $`(t_y,t_z)`$ become smaller as $`\varphi _z`$ decreases, and the energy scale (width of the Landau band $`4(t^{}+t_z)`$ with $`t^{}t_z`$) shrinks for smaller $`\varphi _z`$, so that it will be hard to resolve the butterfly structure. For typical quasi-1D organic conductors such as (TMTSF)<sub>2</sub>X we have $`t_x:t_y:t_z=1:0.1:0.01`$ with $`a,b,c10\mathrm{\AA }`$, and we can estimate the required $`\varphi _z0.1`$, which corresponds to $`B400`$T. It is still huge, but much smaller than the value required for 2D and around the border of experimental feasibility.
While we have discussed appropriate values for given $`(t_y,t_z)`$, are there restrictions on $`(t_y,t_z)`$ to have butterflies? Binding of a well must be so strong that the transfer to second neighbors is negligible. If one approximates a well in $`V^{(1)}`$ with a harmonic potential, the quantized energies are $`(n+\frac{1}{2})\mathrm{}\omega `$. Then the $`n`$-th state is strongly bound to each well when this energy is smaller than $`4t_y`$, the depth of a well. Hence $`t_y`$ should not be too small ($`\sqrt{t_y}>\varphi _z`$), otherwise we have trivial Bragg-reflection gaps only. Also, the residual potential $`V^{(2)}`$ whose amplitude is $`2t_z`$ must be weaker than $`\mathrm{}\omega `$ (i.e., $`t_z<\varphi _z\sqrt{t_y}`$) so that different Landau bands are not mixed. All the conditions (those discussed in this paragraph as well as $`\varphi _z,\varphi _y1,t_y\varphi _zt_z\varphi _y`$) can be interpreted in the semiclassical quantization involving the cross sections of equipotential surfaces, but the essential hopping ($`t^{}`$) between adjacent cross-sectional orbits is outside the scope of the semiclassical picture.
Now we come to our goal of calculating the Hall conductivity when the Fermi level lies in each gap of the 3D butterfly. The mapping does indeed enables us to accomplish this through identifying the topological invariants in the general formula for the Hall conductivity for 3D Bloch electrons by Kohmoto-Halperin-Wu. In the formula the Hall conductivity tensor is expressed as
$$\sigma _{ij}=\frac{e^2}{2\pi h}\underset{k}{}ฯต_{ijk}G_k$$
(10)
when Fermi energy is in a gap. Here $`ฯต_{ijk}`$ is a unit antisymmetric tensor, $`๐ฎ=\mu _1๐^{\mathbf{}}+\mu _2๐^{\mathbf{}}+\mu _3๐^{\mathbf{}}`$ with $`๐^{\mathbf{}},๐^{\mathbf{}},๐^{\mathbf{}}`$ being the primitive reciprocal lattice vectors, and $`\mu _1,\mu _2,\mu _3`$ are topological invariants specifying each gap (i.e., remain constant when we change the direction of $`B`$ in the present context). For an orthogonal lattice we have simply $`\sigma _{yz}=\frac{e^2}{h}\frac{\mu _1}{a},\sigma _{zx}=\frac{e^2}{h}\frac{\mu _2}{b},\sigma _{xy}=\frac{e^2}{h}\frac{\mu _3}{c}`$.
So we have only to determine invariant integers, which are subject to a Diophantine equation,
$$\frac{r}{Q}=\lambda +\frac{P}{Q}n_x\mu _1+\frac{P}{Q}n_y\mu _2+\frac{P}{Q}n_z\mu _3,$$
(11)
where we have assumed a rational magnetic flux, $`(\varphi _x,\varphi _y,\varphi _z)=\frac{P}{Q}(n_x,n_y,n_z)`$ ($`P,Q`$: integers, $`n_x`$ etc have no common divisors), $`r`$ the number of occupied bands, and $`\lambda `$ another topological invariant. Although the solution of the Diophantine equation is not unique, Thouless et al. argued for 2D that there is a restriction on the integers that decides the solution uniquely. Kohmoto et al. conjecture the uniqueness of the solution in 3D in analogy with the 2D case, where the restriction for 3D reads $`|\mu _1n_x+\mu _2n_y+\mu _3n_z|<Q/2`$.
We can then calculate the Hall conductivity for the 3D butterfly. We assume $`(\varphi _x,\varphi _y,\varphi _z)=P/Q(0,n_y,n_z)`$ and $`t_y\varphi _zt_z\varphi _y`$, for which the effective flux in eq.(7) is $`\varphi =\varphi _y/\varphi _z=n_y/n_z`$. Hence each Landau band should split into $`n_z`$ butterfly subbands. Let us consider the situation where $`E_F`$ lies just above the $`m`$th subband in the $`l`$th Landau band from the bottom, i.e., $`(ln_z+m)`$ subbands altogether. Each subband is shown to comprise $`P`$ bands, so that the gap has an index $`r=(ln_z+m)P`$. Substituting this in the Diophantine eq.(11), we have $`(ln_z+m)P=Q\lambda +Pn_y\mu _2+Pn_z\mu _3`$. Since $`P`$ and $`Q`$ have no common divisors, $`s`$ must be a multiple of $`P`$, and with the above restriction one has $`\lambda =0`$, and we end up with $`ln_z+m=n_y\mu _2+n_z\mu _3`$ for the 3D butterfly in the lower half of the entire band. Thus we can determine $`\mu _2,\mu _3`$ for an arbitrary gap in the 3D butterfly as explicitly displayed in Fig.3(a).
If we compare with a corresponding plot for 2D in Figure 3(b), we recognize a beautiful consequence of the 2D-3D mapping established here as a unsuspected one-to-one correspondence between the Hall conductivities on 2D and 3D butterflies as a whole (i.e., for a set of topological invariants attached to the recursive gaps). Namely, the Hall conductivity in 2D is given by $`\sigma _{2\mathrm{D}}=\frac{e^2}{h}t`$, where $`t`$ is an integer in a 2D Diophantine equation $`r=qs+pt`$. If we compare this with the 3D Diophantine equation, $`m=n_z(\mu _3l)+n_y\mu _2`$, the mapping dictates a correspondence $`n_yp,n_zq,mr`$, so that the invariant integers should translate as
$$\mu _3ls,\mu _2t.$$
(12)
This implies that $`\sigma _{zx}`$ in 3D plays the role of $`\sigma _{2\mathrm{D}}`$.
Future problems are to extend the effective theory to the the arbitrary orientation of $`๐ฉ`$, for which Harperโs equation in the arbitrary $`๐ฉ`$ has been discussed in an existing literature . We wish to thank Mahito Kohmoto for discussions.
|
warning/0006/astro-ph0006401.html
|
ar5iv
|
text
|
# Bayesian Methods for Cosmological Parameter Estimation from Cosmic Microwave Background Measurements
## I Introduction
Recent observations of the angular power spectrum of the anisotropy of the cosmic microwave background (CMB) have created much excitement. This data may be used to estimate cosmological parameters. The maximum in the angular power spectrum around the multipole value $`l200`$ is consistent with inflation and a flat universe . However the apparent absence of a peak at $`l400500`$ may imply that we have underestimated the size of such cosmological parameters as the amount of normal baryonic matter or cold dark matter. This data, coupled with recent supernovae observations , seems to be leading us to seriously consider the existence of a cosmological constant. Future observations will provide valuable cosmological information .
The extraction of cosmological parameters from the CMB anisotropy data is a complicated and computer intensive activity. This is especially true as the number of cosmological parameters increases as the modal complexities grow. Current analysis exercises have included up to ten parameters; the scaler quadrupole and gravity wave perturbation normalizations ($`A_s`$ and $`A_t`$), the scaler and tensor power-law indices for primordial perturbations ($`n_s`$ and $`n_t`$), the reionization optical depth $`\tau `$, the spatial curvature $`\mathrm{\Omega }_k`$, the energy densities for baryonic matter ($`\mathrm{\Omega }_b`$), cold dark matter ($`\mathrm{\Omega }_{cdm}`$), neutrinos ($`\mathrm{\Omega }_\nu `$) and the vacuum ($`\mathrm{\Omega }_\mathrm{\Lambda }`$) . Other logical cosmological parameters could include the Hubble constant, number of neutrino families and their masses, etc.
Parameter estimation can be comprehensively described within the language of Bayesian statistics. Application of Bayesโ theorem is well suited to astrophysical observations . In Bayesian data analysis the model consists of a joint distribution over all unobserved (parameters) and observed (data) quantities. One conditions on the data to obtain the posterior distribution of the parameters. The starting point of the Bayesian approach to statistical inference is setting up a full probability model that consists of the joint probability distribution of all observables, denoted by $`๐ณ=(z_1,\mathrm{},z_n)`$ and unobservable quantities, denoted by $`๐ฝ=(\theta _1,\mathrm{},\theta _d)`$. Using the notion of conditional probability, this joint PDF $`f(๐ณ,๐ฝ)`$ can be decomposed into the product of the PDF of all unobservables, $`f(๐ฝ)`$, referred to as the prior PDF of $`๐ฝ`$, and the conditional PDF of the observables given the unobservables, $`f(๐ณ|๐ฝ)`$, referred to as the sampling distribution or likelihood, i.e.
$`f(๐ณ,๐ฝ)=f(๐ฝ)f(๐ณ|๐ฝ).`$
The prior PDF contains all the information about the unobservables that is known from substantive knowledge and expert opinion before observing the data. All the information about the $`๐ฝ`$ that stems from the experiment is contained in the likelihood. In the light of the data, the Bayesian paradigm then updates the prior knowledge about $`๐ฝ`$, $`f(๐ฝ)`$, to the posterior PDF of $`๐ฝ`$, $`f(๐ฝ|๐ณ)`$. This is done via an application of Bayes theorem through conditioning on the observations
$`f(๐ฝ|๐ณ)={\displaystyle \frac{f(๐ฝ,๐ณ)}{m(๐ณ)}}f(๐ฝ)f(๐ณ|๐ฝ)`$
where $`m(๐ณ)=f(๐ณ|๐ฝ)f(๐ฝ)๐๐ฝ`$ is the marginal PDF of $`๐ณ`$ which can be regarded as a normalizing constant as it is independent of $`๐ฝ`$. The Bayesian approach is based on the likelihood function but treats the parameters as random variables and assumes a joint prior distribution that summarizes the available information about the parameters before observing the data. In the light of the observations, the information about the unknown parameters is then updated via Bayes theorem to the posterior distribution which is proportional to the product of likelihood and prior density .
The problem of marginalization of parameters has been the Achilles heel of Bayesian approaches to parameter estimation. The inevitable difficulty in calculating the multidimensional integrals necessary for the determination of posterior probability distributions has hindered efforts. However, these impediments have been overcome by the progress made within the last decade in Bayesian computational technology via Markov chain Monte Carlo (MCMC) methods . Since its initial application in digital signal analysis MCMC methods have revolutionized many areas of applied statistics and should have an impact for cosmological parameter estimation from CMB measurements.
In this paper we do NOT implement a MCMC method for estimating cosmological parameters. This task is best left to the active scientists in the field. Instead we describe how researchers may adapt their existing specialized software in order to produce probability distributions functions for the parameters that have been derived in a statistically rigorous manner.
In section II we review methods that have been used to estimate cosmological parameters from CMB measurements, plus techniques for calculating likelihoods. In section III we describe the Bayesian approach to statistical inference and its implementation via Markov chain Monte Carlo methods. In section IV we describe a methods for applying MCMC methods to cosmological parameter estimation with CMB data. Section V presents our conclusions.
## II Current Statistical Methods with CMB Data
### A Parameter Estimation
Various statistical approaches have attempted to estimate cosmological parameters from the CMB data. The maximum likelihood is an example . Likelihood techniques aim at finding the values of the parameters that maximize the likelihood function. However, the maximum of the likelihood is not necessarily the peak value of the parameterโs marginal distribution (its PDF) or its mean. The maximum likelihood parameter estimates will differ from the Bayes estimators. The maximum likelihood is not always good for parameter estimation; for a given parameter, the maximization depends on whether or not other parameters have been integrated out. A proper application of Bayesโ theorem is attempted in some studies through the inclusion of prior distributions . The typical approach is to select a range for the parameters and limit the maximum likelihood technique to that region in parameter space. Sophisticated frequentist techniques are implemented in order to maximize the likelihood, such as the Levenberg-Marquardt method , or the downhill simplex method .
In the methods described above the problem of maginalizing over parameters is non-trivial. Typically the approach is to marginalize over all parameters except the vacuum energy density ($`\mathrm{\Omega }_\mathrm{\Lambda }`$) and the matter energy density ($`\mathrm{\Omega }_m=\mathrm{\Omega }_b+\mathrm{\Omega }_{cdm}`$). This procedure gets prohibitively difficult as the number of parameters increases; in some cases the amount of computation can be expected to rise exponentially with parameter number. Hence, the techniques that have already been used will be difficult to apply to cosmological models with increasing complexity. Marginalizing over parameters will also become difficult if one no longer assumes that the variables are log-normally distributed, or if the cosmological CMB anisotropy signal is actually non-Gaussian .
Numerical maximization techniques used in maximum likelihood estimation, such as the Levenberg-Marquardt method, are only guaranteed to find a local maximum. Once they reached a local maximum they might get stuck in their search and not reach the global maximum. For this reason statisticians have applied simulating annealing, which is a technique for global optimization. Simulated annealing has been attempted for cosmological parameter estimation . Simulated annealing is related to MCMC as its core component is the Metropolis Hastings algorithm. In this method the parameter space is searched in a random way. A new parameter space point is reached with a probability that depends on the likelihood and an effective temperature term. In the limit where the temperature approaches zero the thermodynamics of this parameter space search finds the system approaching the maximum of the likelihood. Stochastic methods such as these substitute deterministic integration by a statistical estimation problem. Although these methods are applicable to high-dimensional problems, they can be very inefficient in certain situations. The shortcomings of simulating annealing are that there is no guarantee that it will actually find the global maximum in finite time. The efficiency depends very much on specifying a good cooling schedule which involves the arbitrary and skillful choice of various cooling parameters that specify the algorithm and determine its performance.
### B Calculating the Likelihood
Implicit in the problem of parameter estimation is the calculation of the likelihood, namely the probability of obtaining the data given a particular model (and its associated parameters). The sizes of the data sets from CMB experiments are immense, so it is impractical to use all the raw data for producing the likelihood. Instead, radical compression of the data is imperative . Even so, a precise derivation of the likelihood is very difficult . Approximate techniques have been developed that take as input the confidence intervals of the experimental results . Software for calculating the likelihoods, RADPACK is freely available at http://flight.uchicago.edu/knox/radpack.html .
In order to calculate the likelihood it is necessary to have the angular power spectrum of this CMB anisotropy for a specific model (with its associated cosmological parameters). This calculational task is accomplished by code such as CMBfast or CAMB . This code accepts the cosmological parameters as input, and returns the angular power spectrum of the CMB anisotropies, $`C_l`$. These software packages serve as the work-horses of current CMB statistical work. For example, the likelihood has been calculated 30,311,820 times in order to cover a region in a ten-dimensional cosmological parameter space . In other studies the likelihood was evaluated as needed within the calculation .
With a large number of parameters it becomes impractical to marginalize by integration. So when attempts are made to produce constraints in the $`\mathrm{\Omega }_\mathrm{\Lambda }\mathrm{\Omega }_m`$ plane approximate marginalization techniques must be applied. One can estimate the parameters by accepting the values from the maximum likelihood on the parameter grid , and then interpolate between grid points with a cubic spline . Another technique is to maximize the likelihood with respect to all remaining parameters for a fixed point in the $`\mathrm{\Omega }_\mathrm{\Lambda }\mathrm{\Omega }_m`$ plane . These methods are further complicated due to degeneracies among the cosmological parameters.
## III Bayesian Posterior Computation via MCMC
The difficulty with the Bayesian approach to parameter estimation is high-dimensional integration. To calculate the normalizing constant of the joint posterior PDF, for instance, requires $`d`$-dimensional integration. Having obtained the joint posterior PDF of $`๐ฝ`$, the posterior PDF of a single parameter $`\theta _i`$ of interest can be obtained by integrating out all the other components, i.e.
$`f(\theta _i|๐ณ)={\displaystyle \mathrm{}f(๐ฝ|๐ณ)๐\theta _1\mathrm{}๐\theta _{i1}๐\theta _{i+1}\mathrm{}๐\theta _d}.`$
Calculation of the posterior mean of $`\theta _i`$ necessitates a further integration, e.g. $`E[\theta _i|๐ณ]=\theta _if(\theta _i|๐ณ)๐\theta _i`$.
As the joint posterior is too complex to sample from directly, we propose to use a MCMC method . Instead of generating a sequence of independent samples from the joint posterior, in MCMC a Markov chain is constructed, whose equilibrium distribution is just the joint posterior. Thus, after running the Markov chain for a certain โburn-inโ period, one obtains (correlated) samples from the limiting distribution, provided that the Markov chain has reached convergence.
One method for generating a Markov chain is via the Metropolis-Hastings algorithm. The MH algorithm was developed by Metropolis et al. and generalized by Hastings . It is a MCMC method which means that it generates a Markov chain whose equilibrium distribution is just the target density, here the joint posterior PDF $`f(๐ฝ|๐ณ)`$. For ease of notation, let us assume that the target density is some $`f(x)`$, where $`x`$ can be $`d`$-dimensional. Then a Markov chain is specified by constructing a transition kernel $`P(x,A)`$, giving the conditional probability to move from state $`x`$ to a point in $`A`$, the Borel $`\sigma `$-field on $`IR^d`$. As the probability to stay in the current state $`x`$, $`P(x,\{x\})`$, is not necessarily 0, we assume that the transition kernel can be expressed as
$`P(x,dy)=p(x,y)dy+r(x)\delta _x(dy)`$
where $`p(x,x)=0`$ and $`\delta _x(dy)=1`$ if $`xdy`$, 0 otherwise, and $`r(x)=1p(x,y)๐y`$. It is well known that if $`p(x,y)`$ satisfies โreversibilityโ or โdetailed balanceโ, i.e.
$`f(x)p(x,y)=f(y)p(y,x)`$
then $`f(x)`$ is the invariant density of the Markov chain. Together with irreducibility and aperiodicity, this gives a sufficient condition for the convergence of the Markov chain to its stationary distribution, see Tierney for further details.
The MH algorithm shares the concept of a generating density with the well-known simulation technique of rejection sampling. However, the candidate generating density $`q(y|x)`$, $`q(y|x)๐y=1`$, can now depend on the current state $`x`$ of the sampling process, and instead of rigorously accepting or rejecting a new candidate $`y`$, it is accepted with a certain acceptance probability $`\alpha (y|x)`$ also depending on the current state $`x`$, and chosen such that the transition probability $`p(x,y)=q(y|x)\alpha (y|x)`$ satisfies detailed balance. This is met by setting
$`\alpha (y|x))=\mathrm{min}\{{\displaystyle \frac{f(y)q(x|y)}{f(x)q(y|x)}},1\}`$
if $`f(x)q(y|x))>0`$ and $`\alpha (y|x)=1`$ otherwise. Note that irreducibility is guaranteed if $`q(y|x)`$ is positive on the support of $`f`$.
The steps of the MH algorithm are therefore:
| Step 0: | Start with an arbitrary value $`x_0`$ |
| --- | --- |
| Step $`i+1`$: | Generate $`y`$ from $`q(.|x_i)`$ and $`u`$ from $`U(0,1)`$ |
| | If $`u\alpha (y|x_i)`$ set $`x_{i+1}=y`$ (acceptance) |
| | If $`u>\alpha (y|x_i)`$ set $`x_{i+1}=x_i`$ (rejection) |
The MH algorithm does not require the normalization constant of the target density. The outcomes from the MH algorithm can be regarded as a sample from the invariant density only after a certain โburn-inโ period. For issues concerning convergence diagnostics, the reader is referred to Cowles and Carlin . Note that an important special case of the MH algorithm is the โindependence chainโ where $`q(y|x)=q(y)`$, i.e. a new candidate is generated independently of the current state $`x`$.
Various methods to assess convergence, i.e. methods used for establishing whether an MCMC algorithm has converged and whether its output can be regarded as samples from the target distribution of the Markov chain, have been developed and implemented in CODA . CODA is a menu-driven collection of SPLUS functions for analyzing the output of the Markov chain. Besides trace plots and the usual tests for convergence, CODA calculates statistical summaries of the posterior distributions and kernel density estimates.
MCMC techniques have been applied in numerous areas, from science to economics. Applications of state-space modeling in finance, e.g. stochastic volatility models applied to time series of daily exchange rates or returns of stock exchange indices, easily have 1000-5000 parameters and the Gibbs sampler shows slow convergence due to high posterior correlations . Specially tailored MCMC algorithms, like multi-move Gibbs samplers or Metropolis-Hastings algorithms, can markedly improve the speed of convergence .
## IV Applying MCMC Methods to CMB Anisotropy Data
It is possible for researchers to generate probability density functions for cosmological parameters from the CMB anisotropy data. This can be done in a Bayesian fashion with the MCMC serving as a means of conducting a proper marginalization over parameters. The implementation of the MCMC method would be relatively straight forward. Instead of calculating the likelihood at uniform locations in the parameter space , one would let the MCMC do its intelligent walk through the space. Uniform a priori distributions for the parameters seems reasonable, so the MCMC would sample the parameter space defined by them. Since the likelihood function can not be written explicitly in terms of the cosmological parameters, but instead in terms of the CMB anisotropy power spectra terms $`C_l`$, it will be necessary to implement a Metropolis-Hastings MCMC routine.
The Markov chain would commence at a randomly selected position in parameter space $`(\theta _1^{(0)},\mathrm{},\theta _d^{(0)})`$. With the parameter set one would then utilize CMBfast or CAMB to generate a set of CMB anisotropy angular power spectrum components ($`C_l`$),.and a likelihood would then be calculated with CMB anisotropy data . New values for the parameters $`(\theta _1^{(1)},\mathrm{},\theta _d^{(1)})`$ would be selected via sampling from the a priori distributions. However, these values would not necessarily be accepted as new values. First they would be used as input to CMBfast or CAMB to generate a set of CMB anisotropy angular power spectrum components ($`C_l`$), and a likelihood would then be calculated with CMB anisotropy data . The new values would be accepted or rejected according the to following test; a random number, $`u`$, would be generated between 0 and 1. If $`u\mathrm{min}[1,f(๐ณ|\theta _1^{(1)},\mathrm{},\theta _d^{(1)})/f(๐ณ|\theta _1^{(0)},\mathrm{},\theta _d^{(0)})]`$ (where $`f(๐ณ|\theta _1^{(1)},\mathrm{},\theta _d^{(1)})`$ is the likelihood in terms of data $`๐ณ`$ and cosmological parameters $`(\theta _1,\mathrm{},\theta _d)`$) then the new parameters is accepted into the chain, if not the next chain element has values equal to that of the previous state. A new set of parameters would then be randomly sampled from the a priori distributions and the procedure would continue.
The generated chain of parameter values would form the set from which the statistical properties would be derived. After running the Markov chain for a certain โburn-inโ period (in order for the Markov chain to reach convergence) one obtains (correlated) samples from the limiting distribution. This process would continue for a sufficiently long time (as determined by convergence diagnostics ).
After the burn-in the frequency of appearance of parameters would represent the actual posterior density of the parameter. From the posterior density one can then create confidence intervals. Summary statistics are produced from the distribution, such as posterior mean and standard deviation. A cross-correlation matrix is also easily produced; this would be of great importance for quantifying the โdegeneracyโ of $`\mathrm{\Omega }_m`$ and $`\mathrm{\Omega }_m`$ . A distinct advantage of the MCMC approach is that computational time scales linearly with parameter number. Hence, the MCMC approach to cosmological parameter estimation may provide the best strategy when testing complex models with numerous parameters. Also, MCMC methods could work with the exact form of the likelihood, although the approximate forms may prove to be sufficiently adequate .
The above method would constitute the simplest implementation of the Metropolis-Hastings method, that of an independence chain. We have just used an acceptance probability $`\alpha (๐ฝ^{}|๐ฝ)`$, defined by
$`\alpha (๐ฝ^{}|๐ฝ)=\mathrm{min}\{{\displaystyle \frac{f(๐ณ|๐ฝ^{})q(๐ฝ|๐ฝ^{})}{f(๐ณ|๐ฝ)q(๐ฝ^{}|๐ฝ)}},1\}`$
where the generating density $`q(๐ฝ^{}|๐ฝ)`$ is the uniform density over the parameter space and thus, in particular, independent of the current state. By using uniform priors, the posterior PDFs in the acceptance probability calculation reduce to the likelihoods. However, the efficiency of a Metropolis-Hastings algorithm depends crucially on the form of the generating density $`q(๐ฝ^{}|๐ฝ)`$. Just using a uniform distribution that does not even depend on the current state $`\theta `$ is the most simple but probably most inefficient way to accomplish the task. Even with a uniform distribution, the algorithm will be irreducible/aperiodic/reversible and thus the Markov chain will converge towards its stationary distribution.
A slightly better way might be to use a uniform distribution in a neighborhood of the current $`๐ฝ`$. Any prior information could be useful, such as correlations that one could use to specify a multivariate normal centered around the current $`๐ฝ`$ with a covariance matrix that takes said correlations into account. The optimization of the Metropolis-Hastings MCMC strategy will inevitably require experimentation with the generating density $`q(๐ฝ^{}|๐ฝ)`$. While this may require some detailed study, the benefit will be the ability to generate posterior distributions for a large number of cosmological parameters.
## V Discussion
A proper Bayesian approach to parameter estimation allows one to simultaneously estimate both dynamic signals and measurement noise. MCMC methods have demonstrated their importance in Bayesian parameter estimation problems with large numbers of parameters. As cosmological models grow in complexity it will become necessary to use techniques such as those discussed here in order to handle marginalization of parameters. Furthermore, a MCMC approach is not restricted to the assumption of Gaussian noise. A heavy-tailed observation error distribution such as a Student-$`t`$-distribution with large degrees of freedom might be more appropriate to allow for crude measurement errors and ensure that resulting estimates will be robust against additive outliers. Non-Gaussian error distributions are readily incorporated into MCMC calculations. We therefore strongly advocate the Bayesian approach via Metropolis-Hastings sampling in future analyses of cosmological parameter estimation from CMB data.
###### Acknowledgements.
This work was supported by the Royal Society of New Zealand Marsden Fund, the University of Auckland Research Committee, and Carleton College.
|
warning/0006/hep-ph0006297.html
|
ar5iv
|
text
|
# I. Introduction
## I. Introduction
Nowadays, there are no doubts the model of QCD vacuum as the instanton liquid (IL) is the most practical instrument on the chiral scale of QCD. It provides, as the lattice calculations recently confirmed, not only the theoretical background for describing spontaneous chiral symmetry breaking (SCSB) but is mostly powerful in the phenomenology of the QCD vacuum and in the physics of light quarks while considered to propagate by zero modes arising from instantons. The origin of gluon and chiral condensates turns out in this picture easily understandable and both are quantitatively calculated getting very realistic values defined by $`\mathrm{\Lambda }_{QCD}`$ and parameters of instanton and anti-instanton ensemble, for example, $`i\psi ^{}\psi (250MeV)^3`$. Moreover, the scale for dynamical quark masses, $`M350MeV`$, naturally appears and pion decay constant, $`f_\pi 100MeV`$, is then transparently calculated.
Another significant advantage of this approach is that the initial formulation starts basically from the first principles and subsequent approximations being well grounded and reliably controlled are plugged in , . It becomes clear especially in recent years when the impressive progress has been reached in understanding the instanton physics on the lattice . Further we are summarizing several things we have learned thinking of the IL theory and trying to answer the challenging questions.
Let us start on that stage of the IL approach when its generating functional has already been taken as factorized one into two factors
$$๐ต=๐ต_g๐ต_\psi ,$$
where eventually $`๐ต_g`$ provides nontrivial gluon condensate and the fermion part $`๐ต_q`$ is responsible to describe the chiral condensate in instanton medium and its excitations. It is usually supposed the functional integral of $`๐ต_g`$ is saturated by the superposition of the pseudo-particle (PP) fields which are the Euclidean solutions of the Yang-Mills equations called the (anti-)instantons
$$A_\mu (x)=\underset{i=1}{\overset{N}{}}A_\mu (x;\gamma _i).$$
(1)
Here $`A_\mu (x;\gamma _i)`$ denotes the field of a single (anti-)instanton in singular gauge with $`4N_c`$ (for the $`SU(N_c)`$ group) coordinates $`\gamma =(\rho ,z,U)`$ of size $`\rho `$ centred at the coordinate $`z`$ and colour orientation defined by the matrix $`U`$. The nontrivial bloc of corresponding $`N_c\times N_c`$ matrices of PP is a part of potential
$$A_\mu (x;\gamma )=\frac{\overline{\eta }_{a\mu \nu }}{g}\frac{y_\nu }{y^2}\frac{\rho ^2}{y^2+\rho ^2}U^{}\tau _aU,y=xz,a=1,2,3,$$
(2)
where $`\tau _a`$ are the Pauli matrices, $`\eta `$ is the โt Hooft symbol , $`g`$ is the coupling constant and for anti-instanton $`\overline{\eta }\eta `$. For the sake of simplicity we do not introduce the distinct symbols for instanton ($`N_+`$) and anti-instanton ($`N_{}`$) and consider topologically neutral IL with $`N_+=N_{}=N/2`$. Utilizing the variational principle the following estimate of $`๐ต_g`$ was found
$$๐ต_ge^S$$
with the action of IL defined by the following additive functional <sup>1</sup><sup>1</sup>1 In fact, the additive property results from the supposed homogeneity of vacuum wave function in metric space. Eq. (3) looks like a formula of classical physics although it describes the ground state of quantum instanton ensemble. Intuitively clear, this definition will be still valid even when the wave function is nonhomogeneous with the nonuniformity scale essentially exceeding average instanton size or, precisely speaking, being larger (or of the order) than average size of characteristic saturating field configuration. Then each instanton liquid element of such a distinctive size will provide a partial contribution depending on the current state of IL, see next Section.
$$S=๐z๐\rho n(\rho )s(\rho ).$$
(3)
The integration should be performed over the IL volume $`V`$ along with averaging the action per one instanton
$$s(\rho )=\beta (\rho )+5\mathrm{ln}(\mathrm{\Lambda }\rho )\mathrm{ln}\stackrel{~}{\beta }^{2N_c}+\beta \xi ^2\rho ^2๐\rho _1n(\rho _1)\rho _1^2,$$
(4)
weighted with instanton size distribution function
$$n(\rho )=Ce^{s(\rho )}=C\rho ^5\stackrel{~}{\beta }^{2N_c}e^{\beta (\rho )\nu \rho ^2/\overline{\rho ^2}},$$
(5)
$$\nu =\frac{b4}{2},b=\frac{11N_c2N_f}{3},\left(\overline{\rho ^2}\right)^2=\frac{\nu }{\beta \xi ^2n},$$
where $`\overline{\rho ^2}=๐\rho \rho ^2n(\rho )/n,n=๐\rho n(\rho )=N/V`$ and $`N_f`$ is the number of flavours. The constant $`C`$ is defined by the variational maximum principle in the selfconsistent way and $`\beta (\rho )=\frac{8\pi ^2}{g^2}=\mathrm{ln}C_{N_c}b\mathrm{ln}(\mathrm{\Lambda }\rho )(\mathrm{\Lambda }=\mathrm{\Lambda }_{\overline{MS}}=0.92\mathrm{\Lambda }_{P.V.})`$ with constant $`C_{N_c}`$ depending on the renormalization scheme, in particular, here $`C_{N_c}\frac{4.66\mathrm{exp}\left(1.68N_c\right)}{\pi ^2\left(N_c1\right)!\left(N_c2\right)!}`$. The parameters $`\beta =\beta (\overline{\rho })`$ and $`\stackrel{~}{\beta }=\beta +\mathrm{ln}C_{N_c}`$ are fixed at the characteristic scale $`\overline{\rho }`$ (an average instanton size). The constant $`\xi ^2=\frac{27}{4}\frac{N_c}{N_c^21}\pi ^2`$ characterizes, in a sense, the PP interaction and Eqs. (3),(4) and (5) describe the equilibrium state of IL. The minor modification of variational maximum principle (see Appendix) leads to the explicit form of the mean instanton size $`\overline{\rho }\mathrm{\Lambda }=\mathrm{exp}\left\{\frac{2N_c}{2\nu 1}\right\}`$ and, therefore, to the direct definition of the IL parameters unlike the conventional variational principle which allows one to extract those parameters solving numerically the transcendental equation only.
The quark fields are considered to be influenced by the certain stochastic ensemble of PPs, Eq. (1), while calculating the quark determinant
$$๐ต_\psi D\psi ^{}D\psi e^{S(\psi ,\psi ^{},A)}_A.$$
Besides, dealing with dilute IL (small characteristic packing fraction parameter $`n\overline{\rho }^4`$) one neglects the correlations between PPs and utilizes the approximation of $`N_c\mathrm{}`$ where the planar diagrams only survive. In addition the fermion field action is approached by the zero modes what means the quark Green function is considered as the superposition of the free Green function $`S_0=(i\widehat{})^1`$ and fermion zero modes $`\mathrm{\Phi }_\pm (xz)`$ which are the solutions of the Dirac equation $`i(\widehat{D}(A_\pm )+m)\mathrm{\Phi }_\pm =0`$ in the field of (anti-)instanton $`A_\pm `$ centred at $`z`$, i.e.
$$S_\pm =S_0\frac{\mathrm{\Phi }_\pm \mathrm{\Phi }_\pm ^{}}{im},$$
(6)
here $`m`$ is the current quark mass and the quark zero mode possesses the following analytic form
$$\left[\mathrm{\Phi }_\pm (x)\right]_{ic}=\frac{\rho }{\sqrt{2}\pi |x|(x^2+\rho ^2)^{3/2}}\left[\widehat{x}\frac{1\pm \gamma _5}{2}\right]_{ij}\epsilon _{jd}U_{dc}$$
with the colour $`c,d`$ and the Lorentz $`i,j`$ indices and the antisymmetric tensor $`\epsilon `$. In fact, there exists the singular term in the limit $`m0`$ but it is selconsistently fixed by the saddle point calculation of quark determinant $`๐ต_\psi `$. In spite of the fact the exact Green function (at $`m0`$) including the terms of the whole series is well known and, moreover, the Green function of instanton molecule has been also established , the simple zero mode approximation (6) is still the most practical in the concrete evaluations. In particular, at $`N_f=1`$ the quark determinant reads
$$๐ต_\psi D\psi ^{}D\psi \mathrm{exp}\left\{๐x\psi ^{}(x)i\widehat{}\psi (x)\right\}\left(\frac{Y^+}{VM}\right)^{N_+}\left(\frac{Y^{}}{VM}\right)^N_{},$$
(7)
$$Y^\pm =i๐z๐U๐\rho n(\rho )/n๐x๐y\psi ^{}(x)i\widehat{}_x\mathrm{\Phi }_\pm (xz)\mathrm{\Phi }_\pm ^{}(yz)i\widehat{}_y\psi (y),$$
where the factor $`M`$ makes the result dimensionless and is also fixed by the saddle point calculation. Amazingly this relatively crude approximation turns out so fruitful to develop (even quantitatively!) the low energy phenomenology of light quarks. The generating functional beyond the chiral limit was obtained in Ref. .
Thus, the IL approach at the present stage of its development looks very indicative, well theoretically grounded and reasonably adjusted phenomenologically. The proper form of generating functional obtained and its verisimilar parameter dependence indicated provide enough predictive power and justify, hence, the approximations made. It dictates the improvements to be done within the approach but, on the other hand, calculating some corrections has clearly no serious prospect. Taking this message as a guiding one we are going to demonstrate that amplifying the approach with an inverse influence of quarks upon the instanton ensemble which is intuitively small effect leads, however, to rather unobvious important conclusions.
We describe this influence (not getting beyond IL and SCSB approach) as a small variation of instanton liquid parameters $`\delta n`$ and $`\delta \rho `$ around their equilibrium values of $`n`$ and $`\overline{\rho }`$ being in full analogy with the description of chiral condensate excitations. Indeed, the result of nontrivial calculation of the functional integral (treating substantially from physical view point the zero quark modes in the fermion determinant) comes to โencodingโ the IL state with just those two parameters. Moreover, the IL density appears in the approach via the packing fraction parameter $`n\overline{\rho }^4`$ only (clear from dimensional analysis) what means one independent parameter existing in practice. It is just the instanton size. The analysis of the quark and IL interaction is addressed in this paper developing our idea of phononlike excitations of IL resulting from the adiabatic changes of the instanton size. Thus, we will suppose the inverse quark influence should be described by these deformable (anti-)instanton configurations which are the field configurations Eq. (2) characterized by the size $`\rho `$ depending on $`x`$ and $`z`$, i.e. $`\rho \rho (x,z)`$.
The paper is organized as follows: in Section II we discuss the modification of quark determinant when the functional integral is saturated by the deformable modes at the minimal number of flavours. Then in Section III we develop the approximate calculation (tadpole approximation) which is based on the saddle point method and the corresponding iteration procedure. Section IV is devoted to the generalization for the multiflavour picture. The meson excitations of quark condensate and calculation of the Gell-MannโOakesโRenner relation when it is provided by the mechanism of quark current mass generation related to the phononlike excitations of IL are analyzed in Section V. The paper includes also Appendix where the fault finding reader gets a chance to control the explicit formulae of the IL parameters and to improve our calculations if is able.
## II. Supplementing phononlike excitations
Apparently, the gist of what we discuss here could be illuminated in the following way. Saddle point method calculation of the functional integral implies the treatment of the action extremals which are the solutions of classical field equations. For the case in hands the action $`S[A,\psi ^{},\psi ]`$ is constructed including the gluon fields $`A_\mu `$, (anti-)quarks $`\psi ^{},\psi `$ and extremals, which are given by the solutions of the consistent system of the Yang-Mills and Dirac equations. As a trial configuration in the IL theory the superposition of (anti-)instantons which is the approximate solution of the Yang-Mills equations (with no reverse influence of the quark fields) and an external field for the Dirac equation simultaneously is considered. We believe it is reasonable to utilize the deformable (crampled) (anti-)instantons $`A_\pm (x;\gamma (x))`$ as the saturating configurations. They just admit of varying the parameters $`\gamma (x)`$ of the Yang-Mills sector of the initial consistent system in order to describe the influence of quark fields in the appropriate variables for the quark determinant.
Taking the action in the form $`S[A_\pm (x,\gamma (x)),\psi ^{},\psi ]`$ we would receive the corresponding variational equation for the deformation field $`\gamma (x)`$ which approaches most optimally (as to the action extremum) PP at nonzero quark fields. In the field theory, for example, the monopole scattering , the Abrikosov vortices scattering are treated in a similar way. However, for IL we avoid the difficulties which come with solving the variational equations if we consider the long-length wave excitations only with the wave length $`\lambda `$ much larger the characteristic instanton size $`\overline{\rho }`$. Indeed, it looks possible because we are searching the kinetic energy of the deformation fields <sup>2</sup><sup>2</sup>2Then we are allowed to take the slowly changing deformation field beyond the integral while calculating the action of deformed instanton. (the one particle contributions) and consider the pair interaction which develops the contact interaction form being calculated in the adiabatic regime .
Let us remind first of all that deriving Eq. (3) we should average over the instanton positions in a metric space. Clearly, the characteristic size of the domain $`L`$ which has to be taken into account should exceed the mean instanton size $`\overline{\rho }`$. But at the same time it should not be too large because the far ranged elements of IL are not โcausallyโ dependent. The ensemble wave function is expected to be homogeneous (every PP contributes to the functional integral being weighted with a factor proportional to $`1/V,V=L^4`$) on this scale. The characteristic configuration which saturates the functional integral is taken as the superposition Eq. (1) with $`N`$ of PP in the volume $`V`$. It is easy to understand that because of an additivity of the functional Eq. (3) describes properly even non-equilibrium states of IL when the distribution function $`n(\rho )`$ does not coincide with the vacuum one and, moreover, allows us to generalize it for the non-homogeneous liquid when the size of the homogeneity obeys the obvious requirement $`\lambda L>\overline{\rho }`$. Besides, we should deal with the weak (comparing to the instanton $`G_{\mu \nu }`$) fields $`g_{\mu \nu }`$ $`(g_{\mu \nu }G_{\mu \nu })`$ and long-length wave (on the scale of the instanton size $`\overline{\rho }`$) perturbations $`\left|\frac{\rho (x,z)}{x}\right|O(1)`$ when the ensemble saturating the functional integral is very close to the vacuum (instanton) one. Then this smoothness or the adiabatic change of instanton size practically dictates another essential simplification defining everywhere the field in the center of instanton $`\frac{\rho (x,z)}{x}\frac{\rho (x,z)}{x}|_{x=z}`$ as a characteristic deformation field and the exact instanton definition in the singular gauge, $`A_\mu ^a(x,z)=\frac{\overline{\eta }_{a\mu \nu }}{g}\frac{}{x_\nu }\mathrm{ln}\left(1+\frac{\rho ^2}{y^2}\right)`$, leads to the following correction to the potential
$$a_\mu ^a(x)=\mathrm{\Phi }_{\mu \nu }^a(x,z)\frac{\rho }{x_\nu },$$
(8)
where $`\mathrm{\Phi }_{\mu \nu }^a(x,z)=\frac{\overline{\eta }_{a\mu \nu }}{g}\frac{2\rho (x,z)}{y^2+\rho ^2(x,z)}`$. Keeping ourselves within the precision accepted here we could take $`\rho (x,z)\overline{\rho }`$, i.e. $`\mathrm{\Phi }_{\mu \nu }^a(x,z)=\mathrm{\Phi }_{\mu \nu }^a(xz)`$. On the other hand the most general form of the correction to the instanton field influenced by the quarks might be calculated if the gluon Green function in the instanton medium is known
$$a_\mu ^a(x,z)=๐\xi D_{\mu \nu }^{ab}(xz,\xi z)J_\nu ^b(\xi z_\psi ),$$
(9)
where $`J_\nu ^b`$ is the current of external quark source, $`z_\psi `$ belongs to the region of long-length wave disturbation and, at last, $`D_{\mu \nu }^{ab}`$ is the Green function of PP in the instanton medium. In fact, this function is not well defined but, seems, for the case in hands we could develop the selfconsistent way to calculate the regularized Green function. The nonsingular propagator behaviour in the soft momentum region is defined by the mass gap of phononlike excitations. Fortunately, the exact form of the Green function occurs unessential here (we are planning to return to the problem of regularized Green function calculation in the forthcoming publication). In the coordinate space it is peaked around the averaged PP size being in nonperturbative regime and, hence, integral Eq. (9) is appraised to be
$$a_\mu ^a(x,z)F_{\mu \nu }^{ab}(xz)J_\nu ^b(zz_\psi ).$$
(10)
Appealing now to Eq. (8) we are capable to get immediately for $`\rho _\nu `$ the following equation
$$\mathrm{\Phi }_{\mu \nu }^a(xz)\frac{\rho (x,z)}{x_\nu }=F_{\mu \nu }^{ab}(xz)J_\nu ^b(zz_\psi ).$$
In fact, the current $`J`$ might be taken constant in the long-length wave approximation. Then neglecting the gradients the following change appears to be justified everywhere $`\left|\frac{\rho (x,z)}{x}\right|\left|\frac{\rho \left(z\right)}{z}\right|`$ since there is no other fields in the problem at all (in the adiabatic approximation). The deformable mode contribution to the functional integral (when the corrections coming from the deformation fields of PP are absorbed) may be estimated as
$$S๐z๐\rho n(\rho )\left\{\frac{\kappa }{2}\left(\frac{\rho }{z}\right)^2+s(\rho )\right\},$$
where $`\kappa `$ is the kinetic coefficient being derived within the quasiclassical approach. Our estimate of it gives the value of a few instanton actions $`\kappa c\beta `$ with the coefficient $`c1.5รท6`$ depending quantitatively on the ansatz supposed for the saturating configurations. Although this estimate is not much meaningful because there is no the vital $`\kappa `$ dependence eventually (becomes shortly clear). Thus, this coefficient should be fixed on a characteristic scale, for example $`\kappa \kappa (\overline{\rho })`$ if we are planning not to be beyond the precision peculiar to the approach. Actually, it means adding the small contribution of kinetic energy type to the action per one instanton only. Such a term results from the scalar field of deformations and affects negligibly the pre-exponential factors of the functional integral. In oneโs turn pre-exponential factors do the negligible influence on the kinetic term as well . If we strive for to be within the approximation we should retain the small terms of the second order in deviation from the point of action minimum only $`\frac{ds\left(\rho \right)}{d\rho }|_{\rho =\rho _c}=0`$ supposing approximately
$$s(\rho )s(\overline{\rho })+\frac{s^{(2)}(\overline{\rho })}{2}\phi ^2,$$
(11)
where $`s^{(2)}(\overline{\rho })\frac{d^2s\left(\rho \right)}{d\rho ^2}|_{\rho _c}=\frac{4\nu }{\overline{\rho ^2}}`$ and the scalar field $`\phi =\delta \rho =\rho \rho _c\rho \overline{\rho }`$ is the field of deviations from the equilibrium value of $`\rho _c=\overline{\rho }\left(1\frac{1}{2\nu }\right)^{1/2}\overline{\rho }`$. Consequently, the deformation field is described by the following Lagrangian density
$$=\frac{n\kappa }{2}\left\{\left(\frac{\phi }{z}\right)^2+M^2\phi ^2\right\}$$
with the mass gap of the phononlike excitations
$$M^2=\frac{s^{(2)}(\overline{\rho })}{\kappa }=\frac{4\nu }{\kappa \overline{\rho ^2}}$$
which is estimated for IL with $`N_c=3`$, for example, in the quenched approximation to be
$$M1.21\mathrm{\Lambda }$$
if $`c=4,\overline{\rho }\mathrm{\Lambda }0.37,\beta 17.5,n\mathrm{\Lambda }^40.44`$ (for the details see the tables of Appendix). The deformation fields $`\phi `$ (with corresponding Jacobian) contribute to the generating functional on the same footing as the quark fields and it looks like
$$๐ต_g^{^{}}D\phi \left|\frac{\delta A}{\delta \phi }\right|\mathrm{exp}\left\{\frac{n\kappa }{2}๐z\left[\left(\frac{\phi }{z}\right)^2+M^2\phi ^2\right]\right\}$$
in full analogy with the fields $`\psi ^{},\psi `$ entering with the functional measure $`D\psi ^{},D\psi `$. Actually the Jacobian contribution <sup>3</sup><sup>3</sup>3Generally, the deformation field $`\rho _\nu `$ and integration variable $`a_\mu ^a`$ (8) are related via the rotation matrix: $`\mathrm{\Omega }_{ab}\mathrm{\Phi }_{\mu \nu }^b(xz)\rho _\nu =\stackrel{~}{a}_\mu ^a`$ and in the long-length wave approximation $`\mathrm{\Phi }`$ might be constant $`\mathrm{\Phi }_{\mu \nu }^b(0)`$($`xz`$). With the rotation matrix spanning the colour field $`a_\mu =\mathrm{\Omega }^1\stackrel{~}{a}_\mu `$ on the fixed axis (on the z axis for SU(2) group, for instance) we can conclude that the vectors $`a_\mu ^z`$ and $`\rho _\nu `$ are, in fact, in one to one correspondence (of course, being within one loop approximation and up to this unessential colour rotation). Thus, the Jacobian occurs an unessential constant. should be omitted in what follows as discussed above.
Analyzing the modifications arising now in the quark determinant $`๐ต_\psi `$ we take into account the variation of fermion zero modes resulting from the instanton size perturbed
$$\mathrm{\Phi }_\pm (xz,\rho )\mathrm{\Phi }_\pm (xz,\rho _c)+\mathrm{\Phi }_\pm ^{(1)}(xz,\rho _c)\delta \rho (x,z),$$
where $`\mathrm{\Phi }_\pm ^{(1)}(u,\rho _c)=\frac{\mathrm{\Phi }_\pm (u,\rho )}{\rho }|_{\rho =\rho _c}`$ and because of the adiabaticity it is valid $`\delta \rho (x,z)`$ $`\delta \rho (z,z)=\phi (z)`$. The additional contributions of scalar fields generate the corresponding corrections in the factors of the kernels $`Y^\pm `$ of Eq. (7) which are treated in the linear approximation in $`\phi `$, i.e.
$`i\widehat{}_x\mathrm{\Phi }_\pm (xz,\rho )\mathrm{\Phi }_\pm ^{}(yz,\rho )i\widehat{}_y\mathrm{\Gamma }_\pm (x,y,z,\rho _c)+\mathrm{\Gamma }_\pm ^{(1)}(x,y,z,\rho _c)\phi (z),`$ (12)
here we introduced the notations
$$\mathrm{\Gamma }_\pm (x,y,z,\rho _c)=i\widehat{}_x\mathrm{\Phi }_\pm (xz,\rho _c)\mathrm{\Phi }_\pm ^{}(yz,\rho _c)i\widehat{}_y,$$
$$\mathrm{\Gamma }_\pm ^{(1)}(x,y,z,\rho _c)=i\widehat{}_x\mathrm{\Phi }_\pm ^{(1)}(xz,\rho _c)\mathrm{\Phi }_\pm ^{}(yz,\rho _c)i\widehat{}_y+i\widehat{}_x\mathrm{\Phi }_\pm (xz,\rho _c)\mathrm{\Phi }_\pm ^{(1)}(yz,\rho _c)i\widehat{}_y$$
(the gradients of scalar field $`\phi `$ are negligible according to the adiabaticity assumption again). It is a simple matter to verify that the right hand side of Eq. (12), being integrated over $`dzdU`$, generates the following kernel (in the momentum space)
$$\frac{1}{N_c}\left[(2\pi )^4\delta (kl)\gamma _0(k,k)+\gamma _1(k,l)\phi (kl)\right]$$
with $`\gamma _0(k,k)=G^2(k),G(k)=2\pi \rho _cF(k\rho _c/2),\gamma _1(k,l)=G(k)G^{}(l)+G^{}(k)G(l),G^{}(k)=\frac{dG\left(k\right)}{d\rho }|_{\rho =\rho _c},F(x)=2x[I_0(x)K_1(x)I_1(x)K_0(x)]2I_1(x)K_1(x)`$, where $`I_i,K_i(i=0,1)`$ are the modified Bessel functions.
In fact, the functional integral of Eq. (7) including the phononlike component may be exponentiated in the momentum space <sup>4</sup><sup>4</sup>4In the metric space we have the nonlocal Lagrangian of the phononlike deformations $`\phi (z)`$ interacting with the quark fields $`\psi ^{},\psi `$, i.e. $``$ $`=`$ $`{\displaystyle ๐x\psi ^{}(x)i\widehat{}_x\psi (x)}{\displaystyle ๐z\frac{n\kappa }{2}\left\{\left(\frac{\phi }{z}\right)^2+M^2\phi ^2(z)\right\}}+`$ $`+`$ $`{\displaystyle \frac{i\lambda _\pm }{N_c}}{\displaystyle ๐x๐y๐z๐U\psi ^{}(x)\{\mathrm{\Gamma }_\pm (x,y,z,\rho _c)+\mathrm{\Gamma }_\pm ^{(1)}(x,y,z,\rho _c)\phi (z)\}\psi (y)}.`$ The physical meaning of the basic phenomenon behind this Lagrangian seems pretty transparent. The propagation of quark fields through the instanton medium is accompanied by the IL disturbance (the analogy with well known polaron problem embarrasses us strongly in this point). with the auxiliary integration over the $`\lambda `$-parameter (see, for example )
$`๐ต_\psi {\displaystyle }{\displaystyle \frac{d\lambda }{2\pi }}\mathrm{exp}\{N\mathrm{ln}\left({\displaystyle \frac{N}{i\lambda VM}}\right)N\}\times `$
$`\times {\displaystyle }D\psi ^{}D\psi \mathrm{exp}\left\{{\displaystyle }{\displaystyle \frac{dkdl}{(2\pi )^8}}\psi ^{}(k)[(2\pi )^4\delta (kl)(\widehat{k}+{\displaystyle \frac{i\lambda }{N_c}}\gamma _0(k,k))+{\displaystyle \frac{i\lambda }{N_c}}\gamma _1(k,l)\phi (kl)]\psi (l)\right\}`$
(we dropped out the factor normalizing to the free Lagrangian everywhere). It is pertinent to mention here the Diakonov-Petrov result comes to the play precisely if the scalar field is switched off.
In order to avoid a lot of the needless coefficients in the further formulae we introduce the dimensionless variables (momenta, masses and vertices)
$$\frac{k\rho _c}{2}k,\frac{M\rho _c}{2}M,\gamma _0\rho _c^2\gamma _0,\gamma _1\rho _c\gamma _1,$$
(13)
the fields in turn
$$\phi (k)(n\kappa )^{1/2}\rho _c^3\phi (k),\psi (k)\rho _c^{5/2}\psi (k),$$
(14)
and eventually for $`\lambda `$ we are using $`\mu =\frac{\lambda \rho _c^3}{2N_c}`$. Then the generating functional takes the following form
$`๐ต{\displaystyle }d\mu ๐ต_g^{^{\prime \prime }}{\displaystyle }D\psi ^{}D\psi D\phi \mathrm{exp}\{N\mathrm{ln}\mu {\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}{\displaystyle \frac{1}{2}}\phi (k)4[k^2+M^2]\phi (k)\}\times `$
(15)
$`\times \mathrm{exp}\left\{{\displaystyle \frac{dkdl}{\pi ^8}\psi ^{}(k)2\left[\pi ^4\delta (kl)(\widehat{k}+i\mu \gamma _0(k,k))+\frac{i\mu }{(n\overline{\rho }^4\kappa )^{1/2}}\gamma _1(k,l)\phi (kl)\right]\psi (l)}\right\},`$
where $`๐ต_g^{^{\prime \prime }}`$ is a part of gluon component of the generating functional which survives after expanding the action per one instanton Eq. (11). The functional obtained describes the IL state influenced by the quarks when all the terms containing the scalar field are collected (see also Appendix). As mentioned above we believe this influence analogous to the reversal impact of phononlike deformations on the quark determinant does not considerably change the numerical results of the IL and SCSB theory. But it is invisible from Eq. (II. Supplementing phononlike excitations) directly how this smallness is reasoned. The scalar field enters the generating functional formally at the same order of the $`\mu `$-expansion as the term providing SCSB (compare the second and third terms of the second exponential function in Eq. (II. Supplementing phononlike excitations)). The suppression arises because the dominant contribution of the Yukawa interaction comes from the quark field condensate which maintains the additional $`\mu `$ smallness. This result prompts, in fact, rather natural scheme of the approximate calculation of the generating functional
$$\psi ^{}\psi \phi =\psi ^{}\psi \phi +\{\psi ^{}\psi \psi ^{}\psi \}\phi .$$
If we believe, for example, the first term ascribes some gluon component of the generating functional then being linearly dependent on the scalar field it produces the small shift from the equilibrium value of the instanton size ($`\rho _c\overline{\rho }`$) $`\phi =\phi ^{}+\delta \phi `$. And the shift $`\delta \phi `$ generates the mass term in the quark sector what means the scheme should be pushed forward to the following expression <sup>5</sup><sup>5</sup>5Let us emphasize the first term is responsible, in a sense, for the quark condensate fluctuations what allows the $`\pi `$ meson to gain its mass. In the second term such fluctuations are suppressed.
$$\psi ^{}\psi \phi =\psi ^{}\psi (\phi ^{}+\delta \phi )+\{\psi ^{}\psi \psi ^{}\psi \}(\phi ^{}+\delta \phi ).$$
Thus, we face the conventional mechanism of the mass generation when it is related to the insignificant variation of the equilibrium instanton size $`\rho _c`$ (or, in usual terms, to falling the scalar field condensate down) produced by the quark condensate. Moreover, it is clear if the variations happen to be of the opposite sign then the field $`\gamma _5\psi `$ (the chiral partner of the field $`\psi `$) should develop the true mass. Before turning to the mechanism of the mass generation below let us calculate the quark determinant integrating formally over the scalar field $`D\phi `$, then find the quark Green function in the tadpole approximation and formulate the equation for the saddle point. It is supposed the phonon component contribution does not affect substantially the results of the SCBS theory. Here we are interested in projecting upon the scale inherent for the scalar field.
## III. Tadpole approximation
The integration leads us to the four fermion interaction and the functional integral can not be calculated exactly. However, due to smallness of scalar field corrections we may find the effective Lagrangian substituting the condensate value in lieu of one of the pairs of quark lines (see Fig. 1.)
$$\psi ^{}(k)\psi (l)\psi ^{}(k)\psi (l)=\pi ^4\delta (kl)TrS(k).$$
In such an approach the diagram with four fermion lines in the lowest order of the perturbation theory in $`\mu `$ is reduced to the two-legs diagram with one tadpole contribution (there are two such contributions because of two possible ways of pairing)
$`{\displaystyle \frac{4(i\mu )^2}{n\overline{\rho }^4\kappa }}{\displaystyle \frac{dkdldk^{}dl^{}}{\pi ^{16}}\gamma _1(k,l)\gamma _1(k^{},l^{})\psi ^{}(k)\psi (l)\psi ^{}(k^{})\psi (l^{})\phi (kl)\phi (k^{}l^{})}`$
$`{\displaystyle \frac{4\mu ^2}{n\overline{\rho }^4\kappa }}{\displaystyle \frac{dk}{\pi ^4}\gamma _1(k,k)\psi ^{}(k)\psi (k)\frac{dl}{\pi ^4}\gamma _1(l,l)TrS(l)D(0)},`$
where the natural pairing definition was introduced
$$\phi (k)\phi (l)=\pi ^4\delta (kl)D(k),D(k)=\frac{1}{4(k^2+M^2)}.$$
It is obvious the factors surrounding $`\psi ^{}(k)\psi (k)`$ has a meaning of quark mass
$$m_f(k)=\frac{\mu }{(n\overline{\rho }^4\kappa )^{1/2}}\gamma _1(k,k)\frac{(2i\mu )}{(n\overline{\rho }^4\kappa )^{1/2}}\frac{dl}{\pi ^4}\gamma _1(l,l)TrS(l)D(0),$$
(16)
(the initial mass term contains the factor $`2`$ when the dimensionless variables are utilized, i.e. takes a form $`2im_f`$.) We are treating it as the current quark mass since the further calculations show its magnitude gets just within the scale interval commonly accepted for the current quark mass and, moreover, it appears in the Gell-MannโOakesโRenner relation.
The contribution of the graph with all the quark lines paired (see Fig. 2)
should be taken into account at the same order of the $`\mu `$ expansion while calculating the saddle point equation
$$\frac{2\mu ^2}{n\overline{\rho }^4\kappa }\left[\frac{dk}{\pi ^4}\gamma _1(k,k)TrS(k)\right]^2\pi ^4\delta (0)D(0)=\frac{1}{2}\frac{\mu ^2}{n\overline{\rho }^4\nu }\frac{V}{\overline{\rho }^4}\left[\frac{dk}{\pi ^4}\gamma _1(k,k)TrS(k)\right]^2.$$
Here we used the natural regularization of the $`\delta `$-function $`\delta (0)=\frac{1}{\pi ^4}\frac{V}{\overline{\rho }^4}`$ in the dimensionless units. Then the quark determinant after integrating over the scalar fields reads
$`๐ต{\displaystyle ๐\mu D\psi ^{}D\psi \mathrm{exp}\left\{N\mathrm{ln}\mu +\frac{2N_c^2}{n\overline{\rho }^4\nu }\frac{V}{\overline{\rho }^4}\mu ^4c^2(\mu )+\frac{dk}{\pi ^4}\psi ^{}(k)2[\widehat{k}+i\mathrm{\Gamma }(k)]\psi (k)\right\}}=`$
(17)
$`={\displaystyle ๐\mu \mathrm{exp}\left\{N\mathrm{ln}\mu +\frac{2N_c^2}{n\overline{\rho }^4\nu }\frac{V}{\overline{\rho }^4}\mu ^4c^2(\mu )+\frac{V}{\overline{\rho }^4}\frac{dk}{\pi ^4}Tr\mathrm{ln}[\widehat{k}+i\mathrm{\Gamma }(k)]\right\}},`$
where the vertex function is defined as
$$\mathrm{\Gamma }(k)=\mu \gamma _0(k,k)+m_f(k),$$
and we introduced the function $`c(\mu )`$ convenient for the practical calculations
$$c(\mu )=\frac{i}{2\mu N_c}\frac{dk}{\pi ^4}\gamma _1(k,k)TrS(k).$$
As it is clear from Eq. (III. Tadpole approximation) the Green function of the quark field is self-consistently defined by the following equation
$$2[\widehat{k}+i\mathrm{\Gamma }(k)]S(k)=1.$$
Searching the solution in the form
$$S(k)=A(k)\widehat{k}+iB(k),$$
we get
$$A(k)=\frac{1}{2}\frac{1}{k^2+\mathrm{\Gamma }^2(k)},B(k)=\frac{1}{2}\frac{\mathrm{\Gamma }(k)}{k^2+\mathrm{\Gamma }^2(k)}.$$
Using Eq. (16) and the definitions of $`\mathrm{\Gamma }(k)`$ and $`B(k)`$ we have the complete integral equation
$$\mathrm{\Gamma }(k)=\mu \gamma _0(k,k)+\frac{N_c}{n\overline{\rho }^4\nu }\mu ^2\gamma _1(k,k)\frac{dl}{\pi ^4}\gamma _1(l,l)\frac{\mathrm{\Gamma }(l)}{l^2+\mathrm{\Gamma }^2(l)},$$
which drives to have the convenient representation of the solution
$$\mathrm{\Gamma }(k)=\mu \gamma _0(k,k)+\frac{N_c}{n\overline{\rho }^4\nu }\mu ^3c(\mu )\gamma _1(k,k).$$
What concerns the function $`c(\mu )`$ it is not a great deal to obtain
$$c(\mu )=\frac{1}{\mu }\frac{dk}{\pi ^4}\gamma _1(k,k)\frac{\mathrm{\Gamma }(k)}{k^2+\mathrm{\Gamma }^2(k)},$$
and, therefore, the complete integral equation for the function <sup>6</sup><sup>6</sup>6 The solution is unique within the $`\mu `$ interval of our interest. It is curious to notice when $`\mu `$ is larger than $`410^2`$ several solution branches for $`c(\mu )`$ emerge (see Fig. 4). Surely, it could be interesting to clarify if there is any correspondence between these branches and the saddle point equation resulting from Eq. (III. Tadpole approximation). However, it is clear wittingly those objects should be heavier than the scale of several hundred $`MeV`$ characteristic for SCSB. It is quite possible such solutions might be associated with some โheavyโ particles if they do exist. $`c(\mu )`$ which is shown in Fig. 3 for $`N_f=1`$. Let us underline the $`N_f`$-dependence of the $`c(\mu )`$ function in the interval of $`\mu `$ determined by saddle point value is unessential. Then we easily obtain for the current quark mass
$$m_f(k)=\frac{N_c}{n\overline{\rho }^4}\frac{\mu ^3}{\nu }c(\mu )\gamma _1(k,k),$$
(18)
and see the cancellation of the kinetic coefficient $`\kappa `$ in $`m_f`$. Thus, it means the precise value of the coefficient is unessential as declared.
We have the following equation for the saddle point of the functional of Eq. (III. Tadpole approximation)
$$\frac{2N_c}{n\overline{\rho }^4}\frac{dk}{\pi ^4}\frac{\mu [\mathrm{\Gamma }^2(k)]_\mu ^{}}{k^2+\mathrm{\Gamma }^2(k)}+\frac{2N_c^2}{n\overline{\rho }^4\nu }\frac{\mu }{n\overline{\rho }^4}[\mu ^4c^2(\mu )]_\mu ^{}=1,$$
(19)
where the prime is attributed to the differentiation in $`\mu `$.
It results from assuming the stationary IL parameters, what is not accurate. We should include another effect produced by the shift of equilibrium instanton size thanks to the quark condensate presence. The modification of the IL parameters ($`n(\mu ),\mathrm{}`$) caused by the Yukawa interaction comes from simple tadpole graph of Fig. 5
in the leading order
$`{\displaystyle \frac{2i\mu }{(n\overline{\rho }^4\kappa )^{1/2}}}{\displaystyle \frac{dkdl}{\pi ^8}\gamma _1(k,l)(\pi ^4)\delta (kl)TrS(k)\phi (kl)}=\mathrm{\Delta }\phi (0),`$
(20)
$`\mathrm{\Delta }={\displaystyle \frac{2i\mu }{(n\overline{\rho }^4\kappa )^{1/2}}}{\displaystyle \frac{dk}{\pi ^4}\gamma _1(k,k)TrS(k)}={\displaystyle \frac{4N_c}{(n\overline{\rho }^4\kappa )^{1/2}}}\mu ^2c(\mu ),`$
and this term generates the small shift of the equilibrium instanton size $`\rho _c\overline{\rho }`$ (remind here $`\phi =\rho \rho _c`$, and $`\phi (0)=๐z\phi (z)`$ is the scalar field in momentum representation) <sup>7</sup><sup>7</sup>7This shift in phononlike component of Lagrangian $`4\frac{M^2}{2}\phi ^2+\mathrm{\Delta }\phi =4\frac{M^2}{2}\phi ^2+\mathrm{}`$ with the definitions $`\phi ^{}=\phi \delta \phi `$ and $`\delta \phi =\frac{\mathrm{\Delta }}{4M^2}`$ generates the mass term in quark sector $`m_f^{^{}}(k)=\frac{\mu }{\left(n\overline{\rho }^4\kappa \right)^{1/2}}\gamma _1(k,k)\delta \phi `$, which has obviously the same form as that of Eq. (18).. The tadpole contribution could be absorbed in making more sophisticated variational procedure of saddle point calculation. Actually, it includes also the variation of the IL parameters as a function of $`\mu `$. But in practice it comes about highly effective to use the simple iterating procedure. At the first step this variation of the IL parameters is not taken into account and the saddle point $`\mu (\rho _c)`$ is calculated from Eq. (19). Then getting new IL parameters (see, Appendix) we have to resolve Eq. (19) again and etc. Five or six iterations are quite enough if we are satisfied with the same precision as in calculating the integrals. In the Table 1 the numerical results (M.S.Z.) are shown for $`N_f=0,1`$ ($`N_f=0`$ corresponds to the quenched approximation for the IL parameters) comparing to those of Diakonov and Petrov (D.P.) where the disturbance of instanton medium was not considered.
Table 1.
| | | D.P. | | | | M.S.Z. | | |
| --- | --- | --- | --- | --- | --- | --- | --- | --- |
| $`N_f`$ | $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | $`N_f`$ | $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | $`m_f`$ |
| $`0`$ | $`5.6810^3`$ | $`341`$ | $`(301)^3`$ | $`0`$ | $`5.5110^3`$ | $`366`$ | $`(297)^3`$ | $`4.29`$ |
| $`1`$ | $`5.1410^3`$ | $`376`$ | $`(356)^3`$ | $`1`$ | $`5.0710^3`$ | $`385`$ | $`(328)^3`$ | $`5.13`$ |
The parameters indicated in the Table 1 are
the dynamical quark mass
$$M(0)=\mathrm{\Gamma }(0)\left(\frac{2}{\rho _c}\right)[MeV],$$
the quark condensate
$$i\psi ^{}\psi =iTrS(x)|_{x=0}=\frac{N_c}{4}\frac{dk}{\pi ^4}\frac{\mathrm{\Gamma }(k)}{k^2+\mathrm{\Gamma }^2(k)}\left(\frac{2}{\rho _c}\right)^3[MeV]^3,$$
and, at last, the current quark mass $`m_f[MeV]`$, which is defined now being mixed with the quark condensate as
$$m_f=\frac{๐km_f(k)TrS(k)}{๐kTrS(k)}.$$
(21)
It is just what is dictated by the Gell-MannโOakesโRenner relation which we calculate below. Through this paper the value of renormalization constant is fixed by $`\mathrm{\Lambda }=280MeV`$. Then the IL parameters are slightly different from their conventional values $`\overline{\rho }(600MeV)^1,\overline{R}(200MeV)^1`$ (see the corresponding tables in Appendix). However, with the minor $`\mathrm{\Lambda }`$ variation the parameters could be optimally fitted. As expected the change of quark condensate is insignificant, the order of several $`MeV`$, what hints the existence of new soft energy scale established by the disturbance which accompanies the quark propagation through the instanton medium.
## IV. Multiflavour approach
In order to match the approach developed to phenomenological estimates we need the generalization for $`N_f>1`$. Then the quark determinant becomes ,
$$๐ต_\psi D\psi ^{}D\psi \mathrm{exp}\left\{๐x\underset{f=1}{\overset{N_f}{}}\psi _f^{}(x)i\widehat{}\psi _f(x)\right\}\left(\frac{Y^+}{VM^{N_f}}\right)^{N_+}\left(\frac{Y^{}}{VM^{N_f}}\right)^N_{},$$
$$Y^\pm =i^{N_f}๐z๐U๐\rho n(\rho )/n\underset{f=1}{\overset{N_f}{}}๐x_f๐y_f\psi _f^{}(x_f)i\widehat{}_{x_f}\mathrm{\Phi }_\pm (x_fz)\mathrm{\Phi }_\pm ^{}(y_fz)i\widehat{}_{y_f}\psi _f(y_f).$$
With phononlike component included every pair of the zero modes $`\mathrm{\Phi }\mathrm{\Phi }^{}`$ acquires the additional term similar to Eq. (12). The appropriate transformation driving the factors $`Y^\pm `$ to their determinant forms is still valid here since the correction term differs from the basic one with the scalar field $`\phi `$. The complete integration over $`dz`$ leads (in the adiabatic approximation $`\phi (x,z)\phi (z)`$) to the transparent Lagrangian form with the momentum conservation of all interacting particles. Besides, we keep the main terms of $`Y^\pm `$ in $`\phi `$ expansion. The quark zero modes generate the factor similar to Eq. (II. Supplementing phononlike excitations) with $`\frac{1}{N_c}`$ being changed by the factor $`\left(\frac{1}{N_c}\right)^{N_f}`$ and then in the leading $`N_c`$ order we have
$`Y^\pm =\left({\displaystyle \frac{1}{N_c}}\right)^{N_f}{\displaystyle ๐z\underset{N_f}{det}\left(iJ^\pm (z)\right)},`$
$`J_{fg}^\pm (z)={\displaystyle \frac{dkdl}{(2\pi )^8}\left[e^{i(kl)z}\gamma _0(k,l)+\frac{dp}{(2\pi )^4}e^{i(kl+p)z}\gamma _1(k,l)\phi (p)\right]\psi _f^{}(k)\frac{1\pm \gamma _5}{2}\psi _g(l)}.`$
While providing the Gaussian form for the functional we perform the integration over the auxiliary parameter $`\lambda `$ together with the bosonization resulting in the integration over the auxiliary matrix $`N_f\times N_f`$ meson fields
$$\mathrm{exp}\left[\lambda det\left(\frac{iJ}{N_c}\right)\right]๐\mathrm{exp}\left\{iTr[J](N_f1)\left(\frac{det[N_c]}{\lambda }\right)^{\frac{1}{N_f1}}\right\}.$$
As a result the generating functional may be written as
$`๐ต={\displaystyle }{\displaystyle \frac{d\lambda }{2\pi }}๐ต_g^{^{\prime \prime }}\mathrm{exp}(N\mathrm{ln}\lambda ){\displaystyle }D\phi \mathrm{exp}\{{\displaystyle }{\displaystyle \frac{dk}{(2\pi )^4}}{\displaystyle \frac{n\kappa }{2}}\phi (k)[k^2+M^2]\phi (k)\}`$
$`{\displaystyle }D_{L,R}\mathrm{exp}{\displaystyle }dz\{(N_f1)[\left({\displaystyle \frac{det[_LN_c]}{\lambda }}\right)^{\frac{1}{N_f1}}+\left({\displaystyle \frac{det[_RN_c]}{\lambda }}\right)^{\frac{1}{N_f1}}]\}`$ (22)
$`{\displaystyle }D\psi ^{}D\psi \mathrm{exp}\{{\displaystyle }{\displaystyle \frac{dk}{(2\pi )^4}}{\displaystyle \underset{f}{}}\psi _f^{}(k)(\widehat{k})\psi _f(k)+i{\displaystyle }dz(Tr[_LJ^+]+Tr[_RJ^{}])\}.`$
Now scalar field interacts with the quarks of the different flavours, nevertheless, the dominant contribution is expected from the tadpole graphs where any pair of the quark fields is taken in the condensate approximation as it happens at $`N_f=1`$
$$\psi _f^{}(k)\psi _g(l)\psi _f^{}(k)\psi _g(l)=\pi ^4\delta _{fg}\delta (kl)TrS(k).$$
As for the condensate itself we obtain it as the nontrivial solution of saddle point equation. For example, it is
$$(_{L,R})_{fg}=\delta _{fg}$$
for the diagonal meson fields. The dimensionless convenient variables (in addition to Eqs. (13), (14)) are the following
$$\frac{}{2}\overline{\rho }^3\mu ,\left(\frac{\lambda \overline{\rho }^4}{(2N_c\overline{\rho })^{N_f}}\right)^{\frac{1}{N_f1}}g.$$
Then the effective action ($`๐ต=๐g๐\mu \mathrm{exp}\{V_{eff}\}`$) in new designations has the form
$$V_{eff}=N(N_f1)\mathrm{ln}g\frac{V}{\overline{\rho }^4}(N_f1)\frac{2\mu ^{\frac{N_f}{N_f1}}}{g}\frac{V}{\overline{\rho }^4}\frac{2N_f^2N_c^2}{n\overline{\rho }^4\nu }\mu ^4c^2(\mu )2N_fN_c\frac{V}{\overline{\rho }^4}\frac{dk}{\pi ^4}\mathrm{ln}\{k^2+\mathrm{\Gamma }^2(k)\},$$
and the saddle point equation reads
$$\frac{2N_c}{n\overline{\rho }^4}\frac{dk}{\pi ^4}\frac{\mu [\mathrm{\Gamma }^2(k)]_\mu ^{}}{k^2+\mathrm{\Gamma }^2(k)}+\frac{2N_fN_c^2}{n\overline{\rho }^4\nu }\frac{\mu }{n\overline{\rho }^4}[\mu ^4c^2(\mu )]_\mu ^{}=1.$$
The quark current mass in Eq. (18) gains the additional factor $`N_f`$ because the scalar nature of the phononlike field requires to match the tadpole quark field condensates of all $`N_f`$ flavours to every vertex
$$m_f(k)=\frac{N_fN_c}{n\overline{\rho }^4}\frac{\mu ^3}{\nu }c(\mu )\gamma _1(k,k).$$
Table 2 complements the Table 1 with the calculations at $`N_f=2`$
Table 2.
| | D.P. | | | | | M.S.Z. | | | | |
| --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- |
| $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | $`f_\pi `$ | $`f_\pi ^{^{}}`$ | $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | $`f_\pi `$ | $`f_\pi ^{^{}}`$ | $`m_f`$ |
| $`4.6610^3`$ | $`422`$ | $`(427)^3`$ | $`135`$ | $`111`$ | $`4.6110^3`$ | $`416`$ | $`(374)^3`$ | $`118`$ | $`97.2`$ | $`12.7`$ |
where $`f_\pi [MeV]`$ is the pion decay constant and
$$f_\pi ^2=\frac{N_cN_f}{8}\frac{dk}{\pi ^4}\frac{\mathrm{\Gamma }^2(k)\frac{k}{2}\mathrm{\Gamma }^{}(k)\mathrm{\Gamma }(k)+\frac{k^2}{4}(\mathrm{\Gamma }^{}(k))^2}{(k^2+\mathrm{\Gamma }^2(k))^2}\left(\frac{2}{\rho _c}\right)^2,$$
$`f_\pi ^{^{}}[MeV]`$ is its approximated form
$$f_\pi ^{}_{}{}^{}2=\frac{N_cN_f}{8}\frac{dk}{\pi ^4}\frac{\mathrm{\Gamma }^2(k)}{(k^2+\mathrm{\Gamma }^2(k))^2}\left(\frac{2}{\rho _c}\right)^2,$$
here $`\mathrm{\Gamma }^{}(k)=\frac{d\mathrm{\Gamma }\left(k\right)}{dk}`$, and the condensate $`i\psi ^{}\psi `$ is implied for the quarks of every flavour.
## V. Meson excitations of the chiral condensate
Here we discuss the meson excitations of the chiral condensate adapting the effective Lagrangian to the suitable variables. The meson degrees of freedom are introduced into the effective Lagrangian (IV. Multiflavour approach) with $`N_f=2`$ by the following substitutions
$`^L=\stackrel{~}{}^L,\stackrel{~}{}^L=(1+\sigma +\eta )UV,U=\mathrm{exp}(i\pi _a\tau ^a),V=\mathrm{exp}(i\sigma _a\tau ^a),`$
$`_R=\stackrel{~}{}^R,\stackrel{~}{}^R=(1+\sigma \eta )VU^{},`$
where $``$ denotes the condensate and $`\sigma ,\eta `$ are the scalar and pseudoscalar meson fields, $`\pi ^a`$ is the isotriplet of the $`\pi `$-mesons, $`\sigma ^a`$ is the vector meson isotriplet. The effective action for the meson field excitations looks then as
$$V_{eff}=\frac{dp}{\pi ^4}\{\pi ^a(p)R_{\pi ^a}(p)\pi ^a(p)+\sigma ^a(p)R_{\sigma ^a}(p)\sigma ^a(p)+\sigma (p)R_\sigma (p)\sigma (p)+\eta (p)R_\eta (p)\eta (p)\},$$
here $`R_{\pi ^a},R_{\sigma ^a},R_\sigma ,R_\eta `$ are the inverse propagators of the corresponding particles. Their exact forms are listed in Ref. but for the $`\pi `$\- and $`\sigma `$-mesons with the phononlike contributions included they are calculated below.
The contributions generated by the quark determinant if the phononlike fields ignored are well investigated in the leading order of the $`N_c`$-expansion. In particular, one of the contributions comes from the diagram A).
It is of the first order in the $`\mu `$-expansion with one quark loop and one external meson leg in which the meson fields are maintained up to the quadratic terms
$`A)2i{\displaystyle }dz{\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}(TrS(k))\mathrm{\Gamma }(k)\{\stackrel{~}{}_{ff}^L(z){\displaystyle \frac{1+\gamma _5}{2}}+\stackrel{~}{}_{ff}^R(z){\displaystyle \frac{1\gamma _5}{2}}\}.`$
Besides, another contribution comes from the diagram B) being of the second order in $`\mu `$ with one quark loop and two external meson legs
$`B)2{\displaystyle }{\displaystyle \frac{dkdl}{\pi ^8}}Tr[S(k)S(l)]\mathrm{\Gamma }(k,l)\mathrm{\Gamma }(l,k)\{\stackrel{~}{}_{fg}^L(lk){\displaystyle \frac{1+\gamma _5}{2}}\stackrel{~}{}_{gf}^L(kl){\displaystyle \frac{1+\gamma _5}{2}}+(L,\gamma _5)(R,\gamma _5)\},`$
where we introduced $`\mathrm{\Gamma }(k,l)=\mu \gamma _0(k,l)+\frac{\mu }{\left(n\overline{\rho }^4\kappa \right)^{1/2}}\gamma _1(k,l)\delta \phi ,\gamma _0(k,l)=G(k)G(l)`$. The determinant terms of the generating functional at $`N_f=2`$ develop the form
$$det\stackrel{~}{}^L+det\stackrel{~}{}^R=1+\sigma ^2+\eta ^2.$$
In addition to this diagrams we have the tadpole diagram C) in which the meson fields are maintaned up to the quadratic terms and which describes the effect of equilibrium instanton size fluctuations when the meson fields are present.
These fluctuations influence the IL parameters <sup>8</sup><sup>8</sup>8 The problems related to the effective chiral Lagrangian (when the phononlike fields included) and its symmetries will be discussed in the separate publication. .
$`C){\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}Tr(S(k))2im_f(k){\displaystyle }dz\{\stackrel{~}{}_{ff}^L(z){\displaystyle \frac{1+\gamma _5}{2}}+\stackrel{~}{}_{ff}^R(z){\displaystyle \frac{1\gamma _5}{2}}\}.`$
In particular, we have for the inverse propagator of the $`\pi `$-meson field
$`R_{\pi ^a}(p)={\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}\{2N_cN_f{\displaystyle \frac{(k,k+p)+\mathrm{\Gamma }(k)\mathrm{\Gamma }(k+p)}{[k^2+\mathrm{\Gamma }^2(k)][(k+p)^2+\mathrm{\Gamma }^2(k+p)]}}\mathrm{\Gamma }(k,k+p)\mathrm{\Gamma }(k+p,k)+`$
$`+2N_cN_f{\displaystyle \frac{\mathrm{\Gamma }^2(k)}{k^2+\mathrm{\Gamma }^2(k)}}+iN_fm_f(k)Tr(S(k))\}.`$
When the pion momentum goes to zero $`p0`$ the first two terms do not generate the massive term because of the quark determinant symmetry. As a result, these terms at small values of $`p`$ originate only term proportional to $`p^2`$ and we limit ourselves with the approximate result for $`\mathrm{\Gamma }=\mu \gamma _0`$ while calculating it. Using the expansion obtained in Ref. we have
$$R_{\pi ^a}(p)=iN_f\frac{dk}{\pi ^4}m_f(k)Tr(S(k))+\beta np^2$$
(23)
with $`\beta n=\frac{N_cN_f}{16}\frac{dk}{\pi ^4}\frac{\mathrm{\Gamma }^2\left(k\right){\scriptscriptstyle \frac{k}{2}}\mathrm{\Gamma }^{}\left(k\right)\mathrm{\Gamma }\left(k\right)+{\scriptscriptstyle \frac{k^2}{4}}\left(\mathrm{\Gamma }^{}\left(k\right)\right)^2}{\left(k^2+\mathrm{\Gamma }^2\left(k\right)\right)^2}\left(\frac{2}{\rho _c}\right)^2`$. Eq. (23) combined with the definition (21) can be given as
$$R_{\pi ^a}(p)=\beta n\left\{\frac{m_fi\psi ^{}\psi N_f}{\beta n}+p^2\right\}.$$
From here we have for the $`\pi `$-meson mass $`m_\pi ^2=\frac{m_fi\psi ^{}\psi N_f}{\beta n}`$ or if the relation between $`\beta `$ and the pion constant $`f_\pi `$ ($`f_\pi ^2=2\beta n`$) plugged in we obtain
$$m_\pi ^2=\frac{2m_fi\psi ^{}\psi N_f}{f_\pi ^2},$$
(24)
what displays Gell-MannโOakesโRenner relation
$$m_\pi ^2=\frac{(m_u+m_d)i[u^{}u+d^{}d]}{f_\pi ^2}.$$
The factor $`2`$ in the numerator of Eq. (24) corresponds just the sum of the $`u`$ and $`d`$ quark masses and $`N_f`$ means the summation of condensates.
The light particle which we introduced and which imitates the scalar glueball properties does not affect significantly the SCSB parameters and correctly describes the soft pion excitations of quark condensate. Meanwhile, the experimental status of this light scalar glueball is very vague. We believe the phononlike excitations could manifest themselves being mixed with the excitations of the quark condensate in the scalar channel. To illuminate the point let us consider the inverse propagator of the $`\sigma `$-meson which is given by the contributions of the diagram $`B)`$ and determinant
$`R_\sigma (p)=2N_cN_f{\displaystyle \frac{dk}{\pi ^4}\frac{(k,k+p)\mathrm{\Gamma }(k)\mathrm{\Gamma }(k+p)}{[k^2+\mathrm{\Gamma }^2(k)][(k+p)^2+\mathrm{\Gamma }^2(k+p)]}\mathrm{\Gamma }(k,k+p)\mathrm{\Gamma }(k+p,k)}+n\overline{\rho }^4.`$
Holding the highest terms of the $`\mu `$-expansion only, when $`\mathrm{\Gamma }=\mu \gamma _0`$, we receive the following result by means of the identity (see, )
$`R_\sigma (p)=N_cN_f{\displaystyle ๐k\frac{[\mathrm{\Gamma }(k)(k+p)_\mu +\mathrm{\Gamma }(k+p)k_\mu ]^2}{[k^2+\mathrm{\Gamma }^2(k)][(k+p)^2+\mathrm{\Gamma }^2(k+p)]}}.`$
In the course of this exercise we have to include the contribution of the โshiftingโ diagram of $`C)`$ type where the field $`\phi `$ (more exactly $`\phi ^{}`$) should be treated as the dynamical one, i.e.
$`C^{})2i\mu {\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}Tr(S(k)){\displaystyle \frac{\gamma _1(k,l)}{(n\overline{\rho }^4)^{1/2}}}{\displaystyle }dz\phi (z)\{\stackrel{~}{}_{ff}^L(z){\displaystyle \frac{1+\gamma _5}{2}}+\stackrel{~}{}_{ff}^R(z){\displaystyle \frac{1\gamma _5}{2}}\}.`$
Then the term of interacting scalar fields which we are interested in looks like $`V_{\phi \sigma }=๐z\mathrm{\Delta }\phi (z)\sigma (z)`$, where $`\mathrm{\Delta }`$ is defined by Eq. (III. Tadpole approximation). At low momenta the diagrams contribute to the effective action (in the dimensionless variables) as
$$2(p^2+M^2)\phi ^22n\overline{\beta }_\sigma (p^2+M_\sigma ^2)\sigma ^2+\mathrm{\Delta }\phi \sigma $$
with $`\overline{\beta }_\sigma `$ standing for the kinetic coefficient of the $`\sigma `$-meson and $`M_\sigma `$ is its mass. Both quantities are derived from the expansion of the inverse propagator in the low momentum region as $`R_\sigma (p)2n\overline{\beta }_\sigma (p^2+M_\sigma ^2)`$. Diagonalizing the quadratic form
$$2(p^2+M^2)\phi ^22(p^2+M_\sigma ^2)\stackrel{~}{\sigma }^2+\stackrel{~}{\mathrm{\Delta }}\phi \stackrel{~}{\sigma },$$
(presented in more adequate variables $`\stackrel{~}{\sigma }=(n\overline{\beta }_\sigma )^{1/2}\sigma ,\stackrel{~}{\mathrm{\Delta }}=\frac{\mathrm{\Delta }}{\left(n\overline{\beta }_\sigma \right)^{1/2}}`$) leads to the following definition of the composite particle masses
$$M_{1,2}^2=\frac{M_\sigma ^2+M^2}{2}\pm \frac{\sqrt{[M_\sigma ^2M^2]^2+\stackrel{~}{\mathrm{\Delta }}^2}}{2}.$$
(25)
It is well known that SCSB is inapplicable when we are going to deal with heavy meson masses. Specifically, it leads to the wrong predictions for the $`\sigma `$-meson because the corresponding mass obtained is of an order $`1/\overline{\rho }1GeV`$ (moreover, the adiabaticity approximation is certainly broken then). The result Eq. (25) aggravates, in a sense, the situation since the hard component leaves for the region of harder masses whereas the light component persists to become lighter. On the other hand, the present observations signal rather the existence of the scalar meson of mass about $`0.5GeV`$ what apparently does not coincide with the light component $`M_2`$. Nevertheless, our consideration having no claims of the quantitative agreement shows the scalar meson could be the mixed particle of pretty large width (similar to the superposition of oscillators with the various fundamental frequencies) describing the excitations of chiral and gluon condensates.
## VI. Conclusion
In this paper we have developed the consistent approach to describe the interaction of quarks with IL. Theoretically, it is based (and justified) on the particular choice of the configurations saturating the functional integral what occurs to be not merely a technical exercise. They are the deformable (crampled) (anti-)instantons with the variable parameters $`\gamma (x)`$ and in the concrete treatment of this paper we play with the variation of the PP size $`\rho (x,z)`$. In a sense, such an ansatz is strongly motivated by the form of quark determinant which is solely dependent on the average instanton size in the SCSB theory. We have demonstrated that in the long-length wave approximation the variational problem of the deformation field optimization turns into the construction of effective Lagrangian for the scalar phononlike $`\phi `$ and quark fields with the Yukawa interaction. Physically, it allows us to analyze the inverse influence of quarks on the instanton vacuum. We have pointed out this influence on the IL parameters as negligible. The modification of the SCSB parameters occurs pretty poor as well. In particular, the scale of quark condensate change amounts to a few $`MeV`$ only. Nevertheless, switching on the phononlike excitations of IL leads to several qualitatively new and interesting effects. The propagation of the quark condensate disturbances over IL happens in this approach to be in close analogy with well-known polaron problem. We imply a necessity to take into account the medium feedback while elementary excitations propagating. The intriguing conclusion comes with realizing it leads to generation of current quark masses $`m_f`$ within QCD itself with their values entirely corresponding to the conventional phenomenological results for $`u`$ and $`d`$ quarks. Moreover, the $`\pi `$-meson being massless pseudo-Goldstone particle in the standard SCSB theory acquires its mass obeying the Gell-MannโOakesโRenner relation when the IL deformations enter the game. Besides, it hints that fitting the parameters $`\overline{\rho },n`$ and renormalization constant $`\mathrm{\Lambda }`$ all together with the alteration of $`s(\rho )`$ profile function we might achieve suitable agreement not only in the order of magnitude. The difficulties which confronted us here illuminate the fundamental problem of gluon field penetration into the vacuum (the instanton vacuum in this particular case) as the most principle one. Indeed, it is a real challenge to answer the question about the strong interaction carrier in the soft momentum region. Perhaps, the light particle of scalar glueball properties which appears inherently in our approach and should manifest itself in the mixture with the excitation of quark condensate in scalar channel ($`\sigma `$-meson) is not bad candidate for that role. By the way, it could be experimentally observed as a wide resonance.
Summarizing, we understand our calculation can not pretend to the precise quantitative agreement with experimental data and see many things to be done. We are planning shortly to consider the problem of instanton profile , to make more realistic description of the PP interaction, to push our ansatz beyond the long-length wave approximation analyzing more precisely โinstanton Jacobianโ $`\left|\frac{\delta A}{\delta \phi }\right|`$.
The authors have benefited from the discussions with many people but especially with N.O. Agasyan, M.M. Musakhanov, Yu.A. Simonov. The paper was accomplished under the Grants of RFFI 97-02-17491 and INTAS (93-0283, 96-0678). Two of us (S.V.M. and A.M.S.) acknowledge Prof. M. Namiki and the HUJUKAI Fund for the permanent financial support.
## Appendix
The contribution of the quark determinant to the IL action is given by the tadpole diagram Eq. (III. Tadpole approximation) which takes the following form when returned to the dimensional variables (see, Eq. (14))
$$\mathrm{\Delta }\phi =\mathrm{\Delta }\frac{(n\kappa )^{1/2}}{\overline{\rho }^3}\phi (0)=\mathrm{\Delta }(n\overline{\rho }^4\kappa )^{1/2}๐\rho \frac{n(\rho )}{n}\frac{dz}{\overline{\rho }^4}\frac{\rho (z)\rho _c}{\overline{\rho }}.$$
Then the IL action, Eq. (3), acquires the additional term
$$S=๐zn\left\{s\mathrm{\Delta }^{}\frac{\rho \rho _c}{\overline{\rho }}\right\},$$
where $`\mathrm{\Delta }^{}=\frac{4N_c}{n\overline{\rho }^4}\mu ^2c(\mu )`$ and the mean action per one instanton is given by the following functional $`s_1=๐\rho s_1(\rho )n(\rho )/n`$ with
$$s_1(\rho )=\beta (\rho )+5\mathrm{ln}(\mathrm{\Lambda }\rho )\mathrm{ln}\stackrel{~}{\beta }^{2N_c}+\beta \xi ^2\rho ^2n\overline{\rho ^2}\mathrm{\Delta }^{}(\rho \rho _c)/\overline{\rho }.$$
In order to evaluate the equilibrium parameters of IL we treat the maximum principle
$$e^Se^{S_0}e^{SS_0}$$
adapting it to the simplest version (when the approximating functional is trivial $`S_0=0`$). In a sense, this choice of the approximating functional should be a little worse than in Ref. . Its only advantage comes from the possibility to get the explicit formulae for the IL parameters in lieu of solving the complicated transcendental equation. In equilibrium the instanton size distribution function $`n(\rho )`$ should be dependent on the IL action only, i.e. $`n(\rho )=Ce^{s(\rho )}`$ where $`C`$ is a certain constant. This argument corresponds to the maximum principle of Ref. . Indeed, if one is going to approach the functional (3) as a local form $`s=๐\rho s_1(\rho )n(\rho )/n`$ where $`s_1(\rho )=\beta (\rho )+5\mathrm{ln}(\mathrm{\Lambda }\rho )\mathrm{ln}\stackrel{~}{\beta }^{2N_c}+\beta \xi ^2\rho ^2n\overline{\rho ^2}`$, it makes the approach self-consistent. The functional difference $`ss_1=๐\rho \{s(\rho )s_1(\rho )\}e^{s(\rho )}/n`$ being varied over $`s(\rho )`$ leads then to the result $`s(\rho )=s_1(\rho )+const`$ keeping into the mind an arbitrary normalization. The maximum principle results in getting the mean action per one instanton as the IL parameters function, for instance, average instanton size $`\overline{\rho }`$. The corrections generated by the โshiftingโ terms turn out to be small and we consider them in the linear approximation in the deviation $`\mathrm{\Delta }`$. The following schematic expansion exhibits how the major contribution appears
$$s_1=\frac{(s+\delta )e^{s\delta }}{e^{s\delta }}\frac{se^s+\delta e^s}{e^s}+\frac{se^s\delta e^ss\delta e^se^s}{e^s^2},$$
(26)
here $`\delta (\mathrm{\Delta })`$ stands for a certain small โshiftingโ contribution and $`s`$ is the action generated by the gluon component only. The last term in Eq. (26) is small comparing to the first one and we neglect it. Then it is clear that evaluating the mean action per one instanton is permissible to hold the gluon condensate contribution $`s`$ only (without the โshiftingโ term $`\delta `$) in the exponential. Hence we have for the mean action per one instanton $`s_1=๐\rho s_1(\rho )n_0(\rho )/n_0`$, and $`n_0(\rho )`$ is the distribution function which does not include the โshiftingโ term <sup>9</sup><sup>9</sup>9The โshiftingโ term changes the mass of phononlike excitation insignificantly. The equilibrium instanton size as dictated by the condition $`\frac{ds\left(\rho \right)}{d\rho }|_{\rho =\rho _c}=0`$ is equal then to $`\rho _c=(\alpha +\mathrm{\Delta }^{}\beta )\overline{\rho },\alpha =\left(1\frac{1}{2\nu }\right)^{1/2},\beta =\frac{1}{4\nu }\left\{1\alpha \frac{\mathrm{\Gamma }\left(\nu +1/2\right)}{\nu ^{1/2}\mathrm{\Gamma }\left(\nu \right)}\right\}`$ and the second derivative of action in the equilibrium point equals to $`s^{^{\prime \prime }}(\rho _c)=\frac{4\nu }{\overline{\rho ^2}}+\frac{2\nu }{\overline{\rho ^2}}\mathrm{\Delta }^{}\left\{\frac{\mathrm{\Gamma }\left(\nu +1/2\right)}{\nu ^{3/2}\mathrm{\Gamma }\left(\nu \right)}\frac{1}{2\nu \alpha }\right\}`$. Another source of corrections to the kinetic coefficient appears while one considers the instanton profile change $`AA+a`$ where the field of correction is $`a\frac{\rho (x,z)}{x}|_{x=z}`$. This mode could appear within the superposition ansatz Eq. (1) and leads to the modifications of quark zero mode ($`D(A+a)\psi =0`$). Fortunately, both corrections to the kinetic term occurs to be numerically small.. It is possible to obtain for the average squared instanton size and the IL density that <sup>10</sup><sup>10</sup>10In order not to overload the formulae with the factors making the results dimensionless, which are proportional to the powers of $`\mathrm{\Lambda }`$, we drop them out hoping it does not lead to the misunderstandings.
$$r^2\overline{\rho ^2}=\nu \left\{1+\frac{\mathrm{\Delta }^{}}{r\overline{\rho }}\frac{\mathrm{\Gamma }(\nu +1/2)}{2\nu \mathrm{\Gamma }(\nu )}\right\}\nu \left\{1+\mathrm{\Delta }^{}\frac{\mathrm{\Gamma }(\nu +1/2)}{2\nu ^{3/2}\mathrm{\Gamma }(\nu )}\right\},$$
(27)
$$n=CC_{N_c}\stackrel{~}{\beta }^{2N_c}\frac{\mathrm{\Gamma }(\nu )}{2r^{2\nu }},$$
(28)
where the parameter $`r^2`$ equals to
$$r^2=\beta \xi ^2n\overline{\rho ^2}.$$
(29)
Expanding $`\mathrm{ln}\rho =\mathrm{ln}\overline{\rho }+\frac{\rho \overline{\rho }}{\overline{\rho }}+\frac{1}{2}\frac{\left(\rho \overline{\rho }\right)^2}{\overline{\rho }^2}+\mathrm{}`$ and using Eq. (27) we can show that
$$\frac{๐\rho n_0(\rho )\mathrm{ln}\rho }{๐\rho n_0(\rho )}=\mathrm{ln}\overline{\rho }+\mathrm{\Phi }_1(\nu ),\frac{๐\rho n_0(\rho )\rho }{๐\rho n_0(\rho )}=\overline{\rho }+\mathrm{\Phi }_2(\nu ),$$
where $`\mathrm{\Phi }_1,\mathrm{\Phi }_2`$ are the certain function of $`\nu `$ independent of $`\overline{\rho }`$. Besides, the average squared instanton size within the precision accepted obeys the equality $`r^2\overline{\rho ^2}=\mathrm{\Phi }(\nu )`$, and $`\mathrm{\Phi }(\nu )`$ is the function of $`\nu `$ only. Then the mean action per one instanton looks like
$$s_1=2N_c\mathrm{ln}\stackrel{~}{\beta }+(2\nu 1)\mathrm{ln}\overline{\rho }+F(\nu )$$
($`F(\nu )`$ is again the function of $`\nu `$ only and its explicit form is unessential for us here). Calculating its maximum in $`\overline{\rho }`$ we receive
$$\overline{\rho }=\mathrm{exp}\left\{\frac{2N_c}{2\nu 1}\right\},\beta =\frac{2bN_c}{2\nu 1}\mathrm{ln}C_{N_c}.$$
From Eqs. (27), (29) we find the IL density to be as
$$n=\nu \frac{e^{\frac{8N_c}{2\nu 1}}}{\beta \xi ^2}\left\{1+\mathrm{\Delta }^{}\frac{\mathrm{\Gamma }(\nu +1/2)}{2\nu ^{3/2}\mathrm{\Gamma }(\nu )}\right\},$$
and handling Eq. (28) we determine the constant $`C`$.
The IL parameters occur to be close to the parameter values of the Diakonov-Petrov approach and are shown in the following
Table 3.
| | | D.P. | | | | M.S.Z. | |
| --- | --- | --- | --- | --- | --- | --- | --- |
| $`N_f`$ | $`\overline{\rho }\mathrm{\Lambda }`$ | $`n/\mathrm{\Lambda }^4`$ | $`\beta `$ | $`N_f`$ | $`\overline{\rho }\mathrm{\Lambda }`$ | $`n/\mathrm{\Lambda }^4`$ | $`\beta `$ |
| 0 | 0.37 | 0.44 | 17.48 | 0 | 0.37 | 0.44 (0.49) | 17.48 |
| 1 | 0.30 | 0.81 | 18.86 | 1 | 0.33 | 0.63 (0.71) | 18.11 |
| 2 | 0.24 | 1.59 | 20.12 | 2 | 0.28 | 1.03 (1.17) | 18.91 |
Here $`N_f`$ is the number of flavours, $`N_c=3`$, and the IL density at $`\mathrm{\Delta }0`$ when the iteration process completed is shown in the parenthesis. It is curious to notice the quark influence on the IL equilibrium state provokes the increase of the IL density.
In Table 4 we demonstrate the mass gap magnitude $`M`$ and the wave length in the โtemporalโ direction $`\lambda _4=M^1`$. To make it more indicative we show also the average distance between PPs from which is clear, in fact, that the adiabatic approximation $`\lambda L\overline{R}>\overline{\rho }`$ is valid for the long-length wave excitations of the $`\pi `$-meson type. All the parameters are taken at $`\kappa =4\beta `$ but the primed ones correspond to the kinetic term value of $`\kappa =6\beta `$.
Table 4.
| $`N_f`$ | $`M\mathrm{\Lambda }^1`$ | $`\lambda \mathrm{\Lambda }`$ | $`M^{^{}}\mathrm{\Lambda }^1`$ | $`\lambda ^{}\mathrm{\Lambda }`$ | $`\overline{R}\mathrm{\Lambda }`$ |
| --- | --- | --- | --- | --- | --- |
| 0 | 1.21 | 0.83 | 0.99 | 1.01 | 1.23 (1.2) |
| 1 | 1.34 | 0.75 | 1.09 | 0.91 | 1.12 (1.09) |
| 2 | 1.45 | 0.69 | 1.18 | 0.84 | 0.99 (0.96) |
Here the parameters in the parenthesis designate the distance between PP ($`\overline{R}=n^{1/4}`$) when the iteration process is completed.
|
warning/0006/cond-mat0006134.html
|
ar5iv
|
text
|
# Symmetric Versus Nonsymmetric Structure of the Phosphorus Vacancy on InP(110)
\[
## Abstract
The atomic and electronic structure of positively charged P vacancies on InP(110) surfaces is determined by combining scanning tunneling microscopy, photoelectron spectroscopy, and density-functional theory calculations. The vacancy exhibits a nonsymmetric rebonded atomic configuration with a charge transfer level $`0.75\pm 0.1`$ eV above the valence band maximum. The scanning tunneling microscopy (STM) images show only a time average of two degenerate geometries, due to a thermal flip motion between the mirror configurations. This leads to an apparently symmetric STM image, although the ground state atomic structure is nonsymmetric. Copyright 2000 by the American Physical Society.
\]
Although it is well known that point defects can exert a profound influence on the physical properties of semiconductors, the determination of the atomic scale geometric and electronic structure of point defects has remained a challenging task for theoretical as well as experimental research. One particularly striking example is anion vacancies on (110) surfaces of III-V semiconductors, where no agreement has been reached regarding even basic properties, such as (i) the symmetry of the atomic structure and (ii) the energy of defect levels in the band gap: Although recent density-functional theory (DFT) calculations for positively charged arsenic vacancies on the GaAs (110) surface agreed that the gallium atoms neighboring the vacancy relax into the surface layer (in contrast with an earlier tight-binding calculation ), one calculation found a rebonded configuration breaking the mirror symmetry of the surface , whereas the other predicted a fully symmetric configuration to have the lowest energy . High resolution scanning tunneling microscopy (STM) images show a density of states preserving the mirror symmetry of the surface at the defect site . Although this seems to favor a symmetric atomic structure, the experimental results could not be matched to results of any of the DFT calculations. Furthermore, scanning tunneling spectroscopy (STS) yielded a local downward band bending of 0.1 eV , whereas surface photovoltage measurements found a band bending of $`0.53\pm 0.3`$ eV at the site of the positively charged As vacancy on $`p`$-doped GaAs(110) surfaces. Concerning the energy levels, the two different DFT calculations predicted the charge transfer levels (+/0) to be 0.32 eV and 0.1 eV , and the lowest Kohn-Sham eigenvalues in the band gap to be 0.73 eV and 0.06 eV above the valence band maximum (VBM). In view of this puzzling situation it is obvious that the interpretation of the experimental STM images as well as the structure of the anion vacancies on (110) surfaces of III-V semiconductors are still under debate .
Discrepancies such as those pointed out above can arise due to limitations of the methods used. On the theoretical side DFT calculations have proven to be powerful to determine the structure of point defects . Yet differences in the size of the supercell, the pseudopotentials employed, or the exchange-correlation functional implemented might lead to different results, such as the defectโs symmetry or the position of the defect levels, especially in situations where two configurations are almost degenerate in energy. On the experimental side the STM images alone do not provide direct information about the atomic structure nor are the experimental conditions, notably the concentration of the vacancies, the same in all experiments. Furthermore the quantities discussed to describe the position of the defect level, i.e., Kohn-Sham eigenvalues within DFT, band bending, and charge transfer levels, are different physical quantities and thus (typically) deviate from each other significantly. In addition, the tip of the STM may affect the structure of the vacancy, since it is well known that the tip can even excite vacancies to migrate . Therefore the combination of state-of-the-art theory and *well defined* experiments is needed to get reliable results and test the correct interpretation of each of the separate results.
In this letter we combine three different methods to determine the geometric and electronic structure of phosphorus (P) vacancies on $`p`$-doped InP(110) surfaces: STM, photoelectron spectroscopy (PES), and DFT. This enables us to determine the charge transfer level of the defect state in the band gap and the equilibrium atomic structure of the vacancy in $`p`$-type material. We identify the rebonded structure as the ground state of the positively charged defect and demonstrate that a thermal flip motion between two mirror configurations results in simulated STM images consistent with the experimental findings. Our results solve the apparent discrepancies of the previous experimental and theoretical studies , provide a comprehensive picture of the vacancy structure, and point out a methodology to determine the structure of defects with high accuracy.
The surface vacancies were produced in high concentrations by annealing initially defect-free $`p`$-doped InP(110) cleavage surfaces \[carrier concentration $`n_{\mathrm{dop}}`$ of $`(1.32.1)\times 10^{18}\mathrm{cm}^3`$\] at $`160^{}\mathrm{C}`$ . Figure 1 shows that this procedure leads indeed to only one type of defect on the surface which was identified as a singly positively charged phosphorus vacancy previously and whose concentration can be controlled by the annealing time . We prepared several samples from the same wafer with a variety of annealing times in an ultrahigh vacuum (UHV) chamber at the โBerliner Elektronenspeicherring fรผr Synchrotronstrahlungโ storage ring for the photoelectron spectroscopy and in another UHV system for the STM measurements ensuring an exact control of defect concentration in both experiments. All measurements were done at room temperature. Angle-resolved PES measurements were used to probe the valence band in normal emission at a photon energy of 50 eV, and the indium $`4d`$ core level. The InP sample was in Ohmic contact with a Au metal surface cleaned by scratching, on which the position of the Fermi level was measured. With the STM we measured constant-current images at a variety of voltages. From these images the vacancy concentrations were determined as a function of the annealing time as well as the spatial distribution of the occupied and empty density of states of the vacancies with atomic resolution.
We first focus on the determination of the energy of the defect state. Here we concentrate on the charge transfer level of the surface vacancy, which is defined as the position of the surface Fermi level at which the defect changes its charge state. This energy level can be calculated by DFT and can be probed by measuring the band bending as shown below. We note that the energy of the charge transfer level might be very different from that of the ionization levels or the Kohn-Sham eigenvalues within DFT. In particular, large deviations between these quantities might occur in systems with large atomic relaxations such as surface and bulk defects .
STM images reveal that during heat treatment of initially defect-free cleavage surfaces the vacancy concentration increases with increasing annealing time. Similarly, the position of the Fermi level, which on well-cleaved defect-free surfaces is close to the top of the valence band (VBM) for $`p`$-type samples, is found to shift toward midgap upon annealing, as measured by PES (see Fig. 2). The filled diamonds (left axis) show the band bending at the surface, as determined through PES, whereas the open circles show the vacancy concentration (from STM, right axis) as a function of the annealing time. The band bending reaches a saturation value of 0.65 eV at vacancy concentrations of $`5\times 10^{12}\mathrm{cm}^2`$. Figure 2 also demonstrates that the Fermi level shift is directly correlated to the increase of the vacancy concentration.
By combining the vacancy concentration and the band bending it induces, it is possible to determine the energy of the charge transfer level in the band gap $`E_{sd}`$, because the charge per surface area in the surface layer $`Q_{ss}`$ is exactly compensated by the charge density per surface area in the depletion layer $`Q_{sc}`$ . The charge per surface area induced by a concentration $`n_{sd}`$ of $`+1e`$ charged P surface vacancies is
$$Q_{ss}=\frac{en_{sd}}{\mathrm{exp}((E_FE_{sd})/kT)+1},$$
(1)
where $`E_F`$ is the Fermi energy. The difference in energy of the charge transfer level and the Fermi level $`E_{sd}E_F`$ is given by $`(E_{sd}E_{sv})(E_vE_{sv})(E_FE_v)`$ (for definitions, see inset in Fig. 2; $`E_v`$ and $`E_{sv}`$ are the positions of the valence band in the bulk and at the surface, respectively) where $`E_vE_{sv}`$ is the band bending $`eV_s`$ measured by photoelectron spectroscopy at the surface. The charge density $`Q_{sc}`$ in the depletion layer compensating $`Q_{ss}`$ is
$$Q_{sc}=\sqrt{2ฯต_0ฯต_rn_{\mathrm{dop}}kT[\mathrm{exp}(\frac{eV_s}{kT})+\frac{eV_s}{kT}1)]}.$$
(2)
The energy difference between the Fermi energy and the valence band maximum in the bulk has been determined for a nondegenerate InP crystal from the carrier concentration according to
$$E_FE_v=kT\mathrm{ln}\left[4n_{\mathrm{dop}}(\frac{2m_pkT}{\pi \mathrm{}^2})^{3/2}\right]$$
(3)
with $`m_p`$ being the mass of the holes.
These equations permit a determination of the difference in energy of the charge transfer level $`E_{sd}`$ and the valence band maximum at the surface $`E_{sv}`$ using the vacancy concentration dependence of the band bending (see Fig. 2). We obtain an energy for the charge transfer level of the P vacancy of $`0.75\pm 0.1`$ eV above the valence band maximum at room temperature. Note that the knowledge of the defect concentrations is crucial to determine the charge transfer level from a measurement of the band bending.
At this stage we compare this energy value with results from our total-energy density-functional theory calculations. Details of the calculational method used are described in Refs. . Here we only increased the cutoff of the plane wave basis set to 15 Ry in the electronic structure calculation to ensure convergence of the indium pseudopotential. We used a $`2\times 4`$ and 6 layers thick supercell, which is large enough here, since the calculated bond lengths changed less than 0.01 ร
and the defect formation energy by less than 0.01 eV when increasing the size of the supercell. For the $`+1e`$ charged P vacancy the calculation yields two different equilibrium atomic configurations, whose total energies differ by only 0.07 eV. The vacancy configuration with the lowest energy exhibits a rebonding of one of the neighboring surface atoms with the indium in the second layer leading to a nonsymmetric atomic structure with respect to the ($`1\overline{1}0`$) mirror plane of the defect-free (110) surface \[see atomic structure superposed on Fig. 3(a)\]. The calculated indium-indium spacing of this configuration is 2.89 ร
as compared to 2.98 ร
for the symmetric ground state structure of the neutral defect. A similar relaxation has also been calculated for the positively charged As vacancy on GaAs(110) . From our calculations we find a charge transfer level (+/0) at 0.52 eV above the VBM for the nonsymmetric vacancy. For the energetically less stable symmetric configuration this level is at 0.45 eV. Both values deviate from the measured charge transfer level by roughly 0.2 eV. We attribute this deviation to the well-known underestimation of energy levels in local density approximation. An approximate upper limit of the error is the difference between the calculated and experimental band gap. Using our calculated value (1.26 eV) and the experimental one (1.42 eV) we get a systematic error of about 0.16 eV. We can therefore conclude that the theoretical and experimental positions of the defect levels agree well within the error margins. We note, however, that the systematic error for the calculated energies of the charge transfer levels is too large to identify the symmetry of the vacancy on the position of the defect level only. In the following we will show that an unambiguous identification can be obtained based on a comparison of the total energies and an analysis of the STM micrographs.
We now focus on the atomic structure. STM images of $`+1e`$ charged P vacancies \[Fig. 3(d)\] always exhibit an apparently symmetric density of states with respect to the ($`1\overline{1}0`$) mirror plane. The STM images thus suggest a symmetric vacancy structure although in our calculations we find the nonsymmetric structure to be lower in energy. In order to clarify this apparent conflict, we have calculated in Fig. 3(a)$``$(c) simulations of occupied (upper frames) and empty (lower frames) state STM images from the local density of states of different vacancy configurations. Frames 3(a1) and 3(a2) show the simulated STM images for the nonsymmetric case, whereas frames 3(b1) and 3(b2) show the same for the symmetric vacancy configuration. Both vacancy configurations lead to clearly distinguishable STM images: the nonsymmetric vacancy exhibits a pronounced asymmetry. The P atom next to the rebonded indium atom \[upper atom in Fig. 3(a)\] has a brighter dangling bond than the P atom at the opposite side of the vacancy and the rebonded indium atom has a weaker maximum in the empty states. Lateral displacements are also visible. These effects are purely electronic, because the calculated height difference is only 0.11 and 0.07 ร
for the cations and anions, respectively. In contrast, the symmetric vacancy shows the expected symmetry with respect to the ($`1\overline{1}0`$) mirror plane \[Fig. 3(b)\]. The images of the occupied states exhibit very little changes of the neighboring occupied dangling bonds \[Fig. 3(b1)\], and the separation of the two neighboring empty dangling bonds is considerably reduced \[Fig. 3(b2)\]. If we compare these simulations with the experimental STM images we find a symmetric density of states \[Fig. 3(d)\] in conflict with the energetically favorable nonsymmetric vacancy configuration in Fig. 3(a). However, also the symmetric vacancy configuration does not agree with the experimental STM images: The empty state STM images do not show any reduced separation of the two neighboring empty dangling bonds in contrast to the simulation of the symmetric vacancy structure \[Fig. 3(b2)\]. Thus none of the static equilibrium vacancy configurations can be matched in detail to the experimental observation.
This conflict can be resolved by closer inspection of the barrier between the two possible mirror configurations of the asymmetric vacancy. We estimated an upper limit for this barrier by mapping the energy along the direct reaction path between the symmetric and nonsymmetric vacancy configurations. On this basis we found an upper limit of $`0.08`$ eV for the barrier between the two nonsymmetric configurations. The low value of the barrier implies that, at room temperature, the vacancy flips between the two nonsymmetric configurations at a rate significantly higher than the time resolution of the STM (1 to 10 Hz). We estimate a flip rate at room temperature of 0.1 THz using the transition state theory, the optical phonon frequency of 2 THz as attempt frequency, and the calculated barrier. This concept of a thermally activated flip motion has been well established for dimers at the Si(001) surface . Based on these findings we interpret the STM image as a time average of the flipping vacancy.
In order to test the model of a thermal flip motion we have performed STM image simulations based on this flip mechanism. In Fig. 3(c) we assumed that the vacancy flips such that the density of states probed by the STM is the average of the rebonded and nonrebonded sides of the vacancy (average of the two mirror configurations). The images agree very well with the experiment. In particular, two major features โ the nearly unchanged separation of the neighboring empty dangling bonds \[Fig. 3(c2)\] and the depression of the neighboring occupied dangling bonds \[Fig. 3(c1)\] โ are well reproduced. Close inspection of Figs. 3(c) and 3(d) clearly demonstrates that the images of the thermally flipping vacancy configuration explains the STM images best. Note that the somewhat higher intensity in the experimental empty state images is due to the local electrostatic potential arising from the positive chare and that the small intensity of empty density of states above the anions decays quickly into the vacuum such that the STM cannot probe it .
In conclusion, we have demonstrated a methodology to determine the atomic structure and the energy of charge transfer levels in the band gap of surface defects, by combining scanning tunneling microscopy, photoelectron spectroscopy, and density-functional theory calculations. From linking STM and PES data we find a charge transfer level (+/0) $`0.75\pm 0.1`$ eV above the VBM for the example of P vacancies on InP(110) in good agreement with DFT calculations. By comparing high resolution STM images with calculations, we determine the vacancy to exhibit a rebonded nonsymmetric structure. The vacancy flips thermally between its two mirror configurations. This localized vibrational mode leads to an apparently symmetric density of states in STM images, although the vacancy has a nonsymmetric atomic structure. Similar measurements may help to understand on the atomic scale the properties of a wide range of materials, where defects play a key role.
We thank Th. Chassรฉ for assistance in an initial experiment, W. Mรถnch for valuable discussions, and K. H. Graf for technical support. This work was supported by the Bundesministerium fรผr Forschung und Technologie under Grant No. 05SE8 OLA7 and by the Deutsche Forschungsgemeinschaft through Sonderforschungsbereich 296 TP A4 and A5 and Grant No. Ne428/2-1.
|
warning/0006/math-ph0006005.html
|
ar5iv
|
text
|
# On the Dynamics of Crystal Electrons, high Momentum Regime
## 1 Introduction
Consider for $`\alpha 0`$ potentials of the form
* $`V:,V(x+\gamma )=V(x)(x,\gamma 2\pi )`$ with
$$V_\alpha ^2:=\underset{n}{}|\widehat{V}(n)|^2(1+n^2)^\alpha <\mathrm{}$$
where $`\widehat{V}(n):=\frac{1}{\sqrt{2\pi }}_0^{2\pi }e^{inx}V(x)๐x`$.
The Stark Wannier Hamiltonian is the selfadjoint operator
$$H^{SW}:=\frac{d^2}{dx^2}x+V$$
defined in $`L^2()`$ by extension from the core $`C_0^{\mathrm{}}()`$; this is a corollary of the Faris Lavine theorem, see .
We shall prove that there are always propagating states and that small potentials with some regularity cannot bind:
###### Theorem 1.1
* Let $`\alpha >0`$ then
$$\sigma _{cont}(H^{SW})\mathrm{};$$
* for $`\alpha >1/2`$ there is a $`c>0`$ such that for $`V_\alpha <c`$
$$\sigma _{pp}(H^{SW})=\mathrm{}.$$
This theorem is a consequence of our main result, Theorem 2.1, which asserts that the probability to be accelerated in the future grows with the momentum of the initial state.
Remark that our results are stated in terms of the StarkโWannier problem but apply as well to the problem of driven quantum rings, see .
The dynamics of crystal electrons described by the present model have been studied since both in mathematics and physics literature, see for a review. The general problem is to understand how the reflections at the band edges accumulate to localize the electron or to create resonances; it is far from being settled. Our contribution is to the question: how do spectral properties change when $`\alpha `$ is diminishing? We refer to for a physical discussion of this theme. Answers for two extreme cases are known: if $`\alpha >5/2`$ then the spectrum of $`H^{SW}`$ is absolutely continuous see and for generalizations; on the other hand there are models (corresponding to $`\alpha <0`$) for which the spectrum has no absolute continuous component, or is even pure point . If $`V`$ is analytic one has certain informations on existence and width of resonances, .
The paper is organized as follows: in the next section we state our main result, Theorem 2.1, precisely and infer Theorem 1.1. The third section contains the dynamical information: the proof of Theorem 2.1 organized in several subsections.
## 2 Spectral properties
Denote the free Stark Hamiltonian by $`H_0^{SW}=\mathrm{\Delta }x`$; by $`D:=i_x`$ the momentum operator which is selfadjoint on $`H^1()`$; $`\chi `$ is the binary fonction with values $`\chi (True)=1,\chi (False)=0`$; $`\chi (Dt[a,b])`$ is the cutoff function in Fourierspace of the interval $`[t+a,t+b]`$; $`cte.`$ a generic constant which may change from line to line;
$$n:=\{\begin{array}{cc}1\hfill & n0\hfill \\ n\hfill & n>0\hfill \end{array}.$$
The main theorem is:
###### Theorem 2.1
Let $`\alpha >0,M>0`$. There exists a $`c=c_{\alpha ,M}>0`$ such that for $`V`$ with $`V_\alpha M`$ and for all $`t,n`$ it holds:
* $`\chi (Dt[n,n+1))(e^{iH^{SW}t}e^{iH_0^{SW}t})`$
$`c_{\alpha ,M}{\displaystyle \frac{V_\alpha }{sign(t)n^{min\{1,\alpha \}}}};`$
* for $`\psi L^2()`$ this implies:
$`\chi (Dt[n,n+1))e^{iH^{SW}t}\psi `$
$`\chi (D[n,n+1))\psi c_{\alpha ,M}{\displaystyle \frac{V_\alpha }{sign(t)n^{min\{1,\alpha \}}}}\psi .`$ (1)
###### Remark 2.2
The dynamical meaning of (ii) is that a large part of a state whose initial momentum is high enough is accelerated, i.e.: For any $`\epsilon >0`$ there is an $`n`$ such that for $`\psi =\chi (D[n,n+1))\psi ,\psi =1`$ it holds
$$\chi (Dt[n,n+1))e^{iH^{SW}t}\psi 1\epsilon (t>0).$$
We show now that the result on the spectrum follows from this:
Proof of Theorem 1.1. For a state $`\psi `$ in the pure point spectral subspace of $`H^{SW}`$ it holds:
$$\underset{R\mathrm{}}{lim}\underset{\pm t0}{sup}\chi (DR)e^{iH^{SW}t}\psi =0,$$
see, for exemple . This implies for $`n`$
$$\underset{t\pm \mathrm{}}{lim}\chi (Dt[n,n+1))e^{iH^{SW}t}\psi =0.$$
(2)
ad (i): Consider the initial state $`\psi =^1\chi (p[n,n+1))`$ where $`^1`$ denotes the inverse unitary Fourier transform. By the inequality (1) it holds for $`t>0,n`$ large enough
$`\chi (Dt[n,n+1))e^{iH^{SW}t}\psi 1c{\displaystyle \frac{V_\alpha }{sign(t)n^{min\{1,\alpha \}}}}{\displaystyle \frac{1}{2}}`$
which contradicts (2). So $`\psi `$ has a part in the continuous subspace of $`H^{SW}`$.
ad (ii): For $`\psi `$ in the pure point subspace take the limits $`tsign(n)\mathrm{}`$ in (1); the equality (2) implies
$$\chi (D[n,n+1))\psi c_{\alpha ,M}\frac{V_\alpha }{|n|^{min\{1,\alpha \}}}\psi (n)$$
which leads to
$$\psi ^2=\underset{n}{}\chi (D[n,n+1))\psi ^2c_{\alpha ,M}V_\alpha ^2\underset{n}{}\frac{1}{|n|^{min\{2,2\alpha \}}}\psi ^2$$
which is a contradiction for $`V_\alpha `$ small enough and $`\alpha >1/2`$; so there are no bound states. $`\mathrm{}`$
## 3 Dynamics
To prove Theorem 2.1 we decompose the operator $`H^{SW}`$ in the Bloch representation. Denote by $`๐`$ the convolution operator $`๐\psi (n)=\frac{1}{\sqrt{2\pi }}_m\widehat{V}(nm)\psi (m)`$ in $`L^2()`$.
$`V`$ is real, so it holds: $`\overline{\widehat{V}(n)}=\widehat{V}(n)`$. We can always substract a constant from $`H^{SW}`$ so we suppose that $`\widehat{V}(0)=0`$.
Consider in
$$L^2([0,1),dk;L^2())_{[0,1)}^{}L^2()๐k$$
the time dependent operator
$$H(t)\psi (k,n)=(n+k+t)^2\psi (k,n)+๐\psi (k,n).$$
$`H_0(t)`$ denotes this operator for $`V=0`$ which has constant domain $`H^2()`$. $`๐`$ is bounded from $`H^2()`$ to $`L^2()`$ for $`\alpha 0`$ so the propagator $`๐`$ generated by $`H(t)`$ is well defined in the strong sense. Its relation to the Wannier Stark propagator is:
$$๐(t)=e^{itx}e^{iH^{SW}t}^1$$
(3)
where $``$ is the Fourier-Bloch transformation
$$\psi (k,n)=\frac{1}{\sqrt{2\pi }}_0^{2\pi }e^{inx}\left\{\underset{\gamma 2\pi }{}e^{ik(x+\gamma )}\psi (x+\gamma )\right\}๐x$$
which maps $`L^2()`$ unitarily onto $`L^2([0,1),dk;L^2())`$.
Remark that $`\psi (k,n)=\psi (k+n)`$. We denote the quantities for the case $`V=0`$ by a subscript $`0`$.
The following statement is a reformulation of Theorem 2.1 in this representation.
Denote $`P_n`$ the projection on the the site $`n`$ in $`L^2()`$ then:
###### Theorem 3.1
For $`\alpha >0,M>0`$ there exists $`c=c_{\alpha ,M}>0`$ such that for $`V`$ with $`V_\alpha M`$ it holds in $`_{[0,1)}^{}L^2()๐k`$ for $`t,n`$:
$`P_n(๐(t)๐_0(t))_{}=`$
$`{\displaystyle _0^t}P_n๐_0^{}๐๐_{}c_{\alpha ,M}{\displaystyle \frac{V_\alpha }{sign(t)n^{min\{1,\alpha \}}}}.`$
We first prove that this is indeed equivalent to Theorem 2.1:
Proof of Theorem 2.1. By the identity (3) it holds:
$`\chi (Dt[n,n+1))(e^{iH^{SW}t}e^{iH_0^{SW}t})_{L^2()}=`$
$`e^{itx}\chi (Dt[n,n+1))e^{itx}^1e^{itx}(e^{iH^{SW}t}e^{iH_0^{SW}t})^1_{}=`$
$`\chi (.+k[n,n+1))(๐(t)๐_0(t))_{}=`$
$`P_n(๐(t)๐_0(t))_{}.`$
so part (i) is equivalent to Theorem 3.1. To see part (ii) observe that
$$e^{iH_0^{SW}t}\chi (Dt[n,n+1))e^{iH_0^{SW}t}=\chi (D[n,n+1))$$
so by (i) and the unitarity of $`e^{iH_0^{SW}t}`$:
$`\chi (Dt[n,n+1))e^{iH^{SW}t}\psi `$
$`\chi (Dt[n,n+1))e^{iH_0^{SW}t}\psi c_{\alpha ,M}{\displaystyle \frac{V_\alpha }{sign(t)n^{min\{1,\alpha \}}}}\psi =`$
$`\chi (D[n,n+1))\psi c_{\alpha ,M}{\displaystyle \frac{V_\alpha }{sign(t)n^{min\{1,\alpha \}}}}\psi .`$
$`\mathrm{}`$
In the rest of the paper we prove Theorem 3.1.
### 3.1 Proof of Theorem 3.1
$`๐(t)=^{}U^k(t)๐k`$. For $`\psi H^2()`$, $`\mathrm{\Omega }^k:=U_{0}^{k}{}_{}{}^{}U^k`$ it holds
$`P_n(U^k(t)U_0^k(t))\psi _{L^2()}=P_n(\mathrm{\Omega }^k(t)๐)\psi `$
$`{\displaystyle _0^t}P_nU_{0}^{k}{}_{}{}^{}๐U_0^k\mathrm{\Omega }^k\psi .`$
Remark that
$$U^k(t,s)=U^0(t+k,s+k)$$
so it is sufficient to estimate
$$_0^tP_nU_{0}^{0}{}_{}{}^{}๐U_0^0\mathrm{\Omega }^0\psi .$$
(4)
We shall give the argument for nonnegative integers $`n`$ and $`t>0`$. For $`n<0,t<0`$ the result then follows from time reversal, i.e.: because for $`๐(t)=T๐(t)T^1`$ with $`T\psi (n):=\overline{\psi }(n)`$ it holds: $`P_n(๐(t)๐_0(t))=P_n(๐(t)๐_0(t))`$. For the remaining cases we shall give an argument later.
In the sequel we shall drop the superscript $`0`$. We shall also suppress the states $`\psi `$; the norm estimates below are to be understood as uniform estimates proven on the dense set $`H^2()`$ and valid by extension in the operator norm.
Rapid oscillations are responsible for the smallness of (4). Remark firstly that with
$$E_n(t):=(n+t)^2\text{ it holds }$$
$$U_0(t)P_n=e^{i_0^tE_n}P_n,$$
secondly that in every time interval
$$I_l:=\frac{l}{2}+[\frac{1}{4},\frac{1}{4})(l)$$
there is exactly one degeneracy, namely
$$E_n\left(\frac{l}{2}\right)=E_{nl}\left(\frac{l}{2}\right)$$
corresponding to some point of stationary phase of $`e^{i_0^t(E_nE_{nl})}`$. Denote
$$_{n,l}:=P_n+P_{nl},_{n,l}^{}:=๐_{n,l}.$$
Heuristically, the contribution
$$_{I_l}P_nU_0^{}๐U_0\mathrm{\Omega }_{I_l}P_nU_0^{}๐_{n,l}U_0\mathrm{\Omega }+_{I_l}P_nU_0^{}๐_{n,l}^{}U_0\mathrm{\Omega }$$
(5)
to (4) is small for two reasons; loosely speaking: the first term describes the probability that a reflection from momentum $`n`$ to $`nl`$ takes place, i.e. that the electron behaves adiabatically. This are less likely for $`n`$ large and $`\widehat{V}`$ small. The second term describes transitions to the other states which is less probable for $`n`$ large because the energetic distance to these states grows like $`|2n+l|`$.
We shall now proceed to the proof according to this intuition. We first treat the reflection to $`nl`$:
###### Lemma 3.2
There is a numerical constant $`c>0`$, independent of $`V`$, such that for all $`\alpha <\beta I_l`$
$$_\alpha ^\beta P_nU_0^{}๐U_0_{n,l}\mathrm{\Omega }c\sqrt{1+V_0}\frac{|\widehat{V}(2n+l)|}{\sqrt{2n+l}}$$
Proof. We supposed that $`\widehat{V}(0)=0`$ so $`P_n๐P_n=0`$; with the notation
$$\phi (t):=_0^t(E_nE_{nl})=_0^t(2n+l)(l+2\tau )๐\tau ;๐_{n,l}:=P_n๐P_{nl};$$
we shall estimate
$$_{I_l}e^{i\phi }๐_{n,l}\mathrm{\Omega }.$$
It will be clear that the reasoning holds uniformly if we integrate only on $`[\alpha ,\beta ]I_l`$. This estimate is done by a stationary phase calculation in the spirit of .
Decompose for an $`a(0,1/2)`$
$$I_l=I_l\{\frac{l}{2}+[\frac{a}{2},\frac{a}{2}]\}\{\frac{l}{2}+[\frac{a}{2},\frac{a}{2}]\}=:I_l^{NS}I_l^S.$$
Then as $`\mathrm{\Omega }=1`$:
$$_{I_l^S}e^{i\phi }๐_{n,l}\mathrm{\Omega }a๐_{n,l}.$$
On the other hand an integration by parts of $`(_te^{i\phi })๐_{n,l}\mathrm{\Omega }/i\dot{\phi }`$ yields
$`{\displaystyle _{I_l^{NS}}}e^{i\phi }๐_{n,l}\mathrm{\Omega }{\displaystyle \frac{๐_{n,l}}{\dot{\phi }}}|_{I_l^{NS}}+{\displaystyle _{I_l^{NS}}}{\displaystyle \frac{๐_{n,l}\dot{\mathrm{\Omega }}}{\dot{\phi }}}+{\displaystyle \frac{\phi ^{\prime \prime }๐_{n,l}\mathrm{\Omega }}{\dot{\phi }^2}}.`$
Now observe that $`|\dot{\phi }(t)|=|2n+l||l+2t||2n+l|a`$, and that $`i\dot{\mathrm{\Omega }}=U_0^{}๐U_0\mathrm{\Omega }`$, so the term on the last expression is smaller than
$`4{\displaystyle \frac{๐_{n,l}}{a|2n+l|}}+{\displaystyle \frac{๐_{n,l}}{|2n+l|}}{\displaystyle _{I_l^{NS}}}\left({\displaystyle \frac{P_{nl}๐}{|l2t|}}+{\displaystyle \frac{|2|}{|l2t|^2}}\right)๐t`$
$`cte.{\displaystyle \frac{๐_{n,l}}{a|2n+l|}}(1+V_0)`$
where we have used that $`P_{nl}๐V_0`$. Thus for $`a(0,1/2)`$:
$$_{I_l}P_nU_0^{}๐U_0_{n,l}\mathrm{\Omega }a๐_{nl}+\frac{1}{a}cte.\frac{๐_{n,l}}{|2n+l|}(1+V_0)$$
The minimum of $`a\alpha +\beta /a`$ for positive $`a`$ is $`2\sqrt{\alpha \beta }`$, thus
$$_{I_l}P_nU_0^{}๐_{n,l}U_0\mathrm{\Omega }c\frac{๐_{n,l}}{\sqrt{|2n+l|}}\sqrt{1+V_0}$$
which implies our assertion as $`๐_{n,l}\frac{1}{\sqrt{2\pi }}|\widehat{V}(2n+l)|`$.
$`\mathrm{}`$
The other levels are separated by large gaps. We start the proof that the transition probability to them is small with a double integration by parts lemma.
We have $`H_0(t)P_n=E_n(t)P_n`$ Denote by
$$\widehat{R}_l(t):=(H_0(t)E_n(t))^1_{n,l}^{}$$
the reduced resolvent. The Friedrichs twiddle operation is fundamental in adiabatic theories. The version needed here is defined for an operator $`A`$ on $`L^2()`$ by
$$\stackrel{~}{A}_l:=P_nA\widehat{R}_l.$$
###### Lemma 3.3
For $`\beta >\alpha `$ it holds on $`H^2()`$
$`{\displaystyle _\alpha ^\beta }U_0^{}P_n๐_{n,l}^{}U_0\mathrm{\Omega }=`$
$`{\displaystyle _\alpha ^\beta }U_0^{}\left(\stackrel{~}{\stackrel{~}{๐}_l๐}_l๐+i\dot{\stackrel{~}{\stackrel{~}{๐}_l๐}_l}\stackrel{~}{๐}_l๐_{n,l}i\dot{\stackrel{~}{๐}}_l\right)U_0\mathrm{\Omega }`$
$`+iU_0^{}\left(\stackrel{~}{๐}_l\stackrel{~}{\stackrel{~}{๐}_l๐}_l\right)U_0\mathrm{\Omega }|_\alpha ^\beta `$
Proof. The twiddle operation is an inverse commutator. A dot $`\dot{}`$ or a prime denotes differentiation. It holds:
$$i_tU_0^{}\stackrel{~}{๐}_lU_0=U_0^{}([\stackrel{~}{๐}_l,H_0E_n]+i\dot{\stackrel{~}{๐}_l})U_0=U_0^{}(P_n๐_{n,l}^{}+i\dot{\stackrel{~}{๐}_l})U_0$$
so an integration by parts and $`i_t\mathrm{\Omega }=U_0^{}๐U_0\mathrm{\Omega }`$ yield
$`{\displaystyle _\alpha ^\beta }U_0^{}P_n๐_{n,l}^{}U_0\mathrm{\Omega }=`$
$`{\displaystyle _\alpha ^\beta }U_0^{}\left(\stackrel{~}{๐}_l๐+i\dot{\stackrel{~}{๐}}_l\right)U_0\mathrm{\Omega }+iU_0^{}\stackrel{~}{๐}_lU_0\mathrm{\Omega }|_\alpha ^\beta .`$
The decomposition $`\stackrel{~}{๐}_l๐=\stackrel{~}{๐}_l๐_{n,l}+\stackrel{~}{๐}_l๐_{n,l}^{}`$, the identity
$$i_tU_0^{}\stackrel{~}{\stackrel{~}{๐}_l๐}_lU_0=U_0^{}(\stackrel{~}{๐}_l๐_{n,l}^{}+i\dot{\stackrel{~}{\stackrel{~}{๐}_l๐}_l})U_0$$
and a second integration by parts imply
$`{\displaystyle _\alpha ^\beta }U_0^{}\stackrel{~}{\stackrel{~}{๐}_l๐}_lP_{n,l}^{}U_0\mathrm{\Omega }=`$
$`{\displaystyle _\alpha ^\beta }U_0^{}\left(\stackrel{~}{\stackrel{~}{๐}_l๐}_l๐+i\dot{\stackrel{~}{\stackrel{~}{๐}_l๐}_l}\right)U_0\mathrm{\Omega }iU_0^{}\stackrel{~}{\stackrel{~}{๐}_l๐}_lU_0\mathrm{\Omega }|_\alpha ^\beta .`$
Thus the assertion is proved. $`\mathrm{}`$
Concerning the second contribution to (5) we shall now proceed to estimate the different terms of
$$_{I_l}P_nU_0^{}๐_{n,l}^{}U_0\mathrm{\Omega }$$
(6)
defined by Lemma 3.3 one after the other.
In the following lemma we collect facts that shall be used frequently and often without comment:
###### Lemma 3.4
Let $`ab`$. For $`\alpha ,\beta >0`$ there is a $`cte.>0`$ such that it holds
$$\underset{j\{a,b\}}{sup}\frac{1}{|ja|^\alpha |jb|^\beta }=\underset{j<\frac{a+b}{2}}{sup}\mathrm{}+\underset{j\frac{a+b}{2}}{sup}\mathrm{}cte.\frac{1}{|ab|^{min\{\alpha ,\beta \}}};$$
and for $`\alpha ,\beta >1`$:
$$\underset{j\{a,b\}}{}\frac{1}{|ja|^\alpha }\frac{1}{|jb|^\beta }\underset{j<\frac{a+b}{2}}{}\mathrm{}+\underset{j\frac{a+b}{2}}{}\mathrm{}cte.\frac{1}{|ab|^{min\{\alpha ,\beta \}}};$$
let $`A`$ be the operator on $`L^2()`$ whose kernel is $`A(i,j)=f(i)g(ij)`$ for some $`f,gL^2()`$, then
$$A=\underset{\phi =\psi =1}{sup}|\phi ,A\psi |f_{L^2()}g_{L^2()}.$$
The smallness of the terms in Lemma (3.3) results from the presence of the reduced resolvent; we shall use that for $`mn,mnl,tI_l`$ it holds as $`|m+n+l|1`$ and so $`|m+n+l+\alpha |\frac{1}{2}|m+n+l|`$ for $`\alpha <\frac{1}{2}`$:
$$\underset{tI_l}{inf}|E_m(t)E_n(t)|=\underset{I_l}{inf}|(mn)(m+n+2t)|\frac{1}{2}|mn||m+n+l|$$
and as in Lemma (3.4):
$$\underset{m\{n,nl\}}{inf}\underset{tI_l}{inf}|E_m(t)E_n(t)|cte.|n(nl)|=cte.|2n+l|.$$
The first relevant term in the integrand of Lemma (3.3) is $`\stackrel{~}{\stackrel{~}{๐}_l๐}_l๐=P_n๐\widehat{R}_l๐\widehat{R}_l๐`$. Now
$$P_n๐(i,j)=\delta _{ni}\widehat{V}(nj);|\widehat{R}_l๐(i,j)|=\chi (in)\chi (inl)\frac{\widehat{V}(ij)}{E_iE_n}.$$
By the third point of Lemma (3.4) we get
$$\widehat{R}_l๐V_0\left(\underset{\{n,nl\}}{}\left|\frac{2}{(in)(i+n+l)}\right|^2\right)^{1/2}cte.V_0\frac{1}{|2n+l|}$$
so
$$\underset{tI_l}{sup}\stackrel{~}{\stackrel{~}{๐}_l๐}_l๐\underset{I_l}{sup}P_n๐\widehat{R}_l๐^2cte.V_0^3\frac{1}{|2n+l|^2}.$$
(7)
For the second term it holds
$$\dot{\stackrel{~}{\stackrel{~}{๐}_l๐}_l}=(P_n๐\widehat{R}_l๐\widehat{R}_l)^{}2๐\dot{\widehat{R}_l}๐\widehat{R}_l$$
so for $`j\{n,nl\}`$:
$$\sqrt{2\pi }๐\dot{\widehat{R}_l}(i,j)=\frac{\widehat{V}(ij)(\dot{E}_j\dot{E}_n)}{(E_jE_n)^2}=\frac{2\widehat{V}(ij)}{(jn)(j+n+2t)^2}$$
so
$$๐\dot{\widehat{R}_l}cte.๐_0\frac{1}{|2n+l|}$$
and together with the estimate of $`๐\widehat{R}_l`$ above it results:
$$\underset{tI_l}{sup}\dot{\stackrel{~}{\stackrel{~}{๐}_l๐}_l}cte.V_0^2\frac{1}{|2n+l|^2}.$$
(8)
Next we discuss $`\dot{\stackrel{~}{๐}}_l=P_n๐\dot{\widehat{R}_l}`$.
$$\sqrt{2\pi }\dot{\stackrel{~}{๐}}_l(i,j)=2\delta _{in}\frac{\widehat{V}(ij)}{(jn)(j+n+2t)^2}\frac{|ij|^\alpha }{|ij|^\alpha }$$
so:
$$\underset{I_l}{sup}\dot{\stackrel{~}{๐}}_lcte.V_\alpha \frac{1}{|2n+l|^{min\{1+\alpha ,2\}}}.$$
(9)
The next contribution to (6) comes from
$$\stackrel{~}{๐}_l๐_{n,l}=P_n๐\widehat{R}_l๐P_n+P_n๐\widehat{R}_l๐P_{nl}.$$
The kernel of the second term is
$$2\pi P_n๐\widehat{R}_l๐P_{nl}(i,j)=\delta _{in}\delta _{j(nl)}\underset{\{n,nl\}}{}\frac{\widehat{V}(nm)\widehat{V}(m+n+l)}{(mn)(m+n+2t)}$$
so
$`\underset{tI_l}{sup}P_n๐\widehat{R}_l๐P_{nl}`$
$`cte.{\displaystyle \underset{\{n,nl\}}{}}\left|{\displaystyle \frac{\widehat{V}(nm)\widehat{V}(m+n+l)}{(mn)(m+n+l)}}\right|{\displaystyle \frac{|mn|^\alpha |m+n+l|^\alpha }{|mn|^\alpha |m+n+l|^\alpha }}`$
$`cte.V_\alpha ^2{\displaystyle \frac{1}{|2n+l|^{1+\alpha }}}`$ (10)
by the Cauchy-Schwartz inequality and Lemma (3.4).
The other term is more difficult, it is the part which is first scattered out of the state $`n`$, then back to it:
$$2\pi P_n๐\widehat{R}_l๐P_n(i,j)=\delta _{in}\delta _{jn}\underset{\{n,nl\}}{}\frac{|\widehat{V}(nm)|^2}{(mn)(m+n+2t)}.$$
Changing coordinates we obtain
$`2\pi P_n๐\widehat{R}_l๐P_n(i,j)=\delta _{in}\delta _{jn}{\displaystyle \underset{p\{0,(2n+l)\}}{}}{\displaystyle \frac{|\widehat{V}(p)|^2}{p(p+2n+2t)}}=`$
$`{\displaystyle \frac{|\widehat{V}(2n+l)|^2}{(2n+l)(4n+l+2t)}}+\delta _{in}\delta _{jn}{\displaystyle \underset{p\{0,\pm (2n+l)\}}{}}{\displaystyle \frac{|\widehat{V}(p)|^2}{p(p+2n+2t)}}=`$
$`{\displaystyle \frac{|\widehat{V}(2n+l)|^2}{(2n+l)(4n+l+2t)}}+\delta _{in}\delta _{jn}{\displaystyle \underset{p\{0,\pm (2n+l)\}}{}}{\displaystyle \frac{|\widehat{V}(p)|^2}{p^2(2n+2t)^2}}`$
because only $`\frac{|\widehat{V}(p)|^2}{p^2(2n+2t)^2}`$, the symmetric part of $`\frac{|\widehat{V}(p)|^2}{p(p+2n+2t)}`$, contributes to the symmetric sum. Physically speaking it is destructive interference of the contributions of the states $`m=n\pm p`$ which is at work here.
Now for $`\alpha >0`$
$`\underset{tI_l}{sup}{\displaystyle \underset{\{0,\pm (2n+l)\}}{}}{\displaystyle \frac{|\widehat{V}(p)|^2}{|p^2(2n+2t)^2|}}{\displaystyle \frac{p^{2\alpha }}{p^{2\alpha }}}`$
$`V_\alpha ^2\underset{p,t}{sup}{\displaystyle \frac{1}{|p^{2\alpha }||p^2(2n+2t)^2|}}V_\alpha ^2{\displaystyle \frac{cte.}{|2n+l|^{min\{1+2\alpha ,2\}}}};`$
to see this recall that for $`p\pm (2n+l)`$
$$\underset{tI_l}{inf}|p^2(2n+t)^2|\frac{1}{4}|p^2(2n+l)^2|,$$
and take the supremum over $`|p|<\frac{2n+l}{2}`$ and $`\pm p\frac{|2n+l|}{2}`$ separately.
Furthermore
$$\frac{|\widehat{V}(2n+l)|^2}{|(2n+l)(2n+l+2n+2t)|}V_0^2\frac{cte.}{|2n+l|^2},$$
so the estimate for the backscattering term is
$`\underset{tI_l}{sup}P_n๐\widehat{R}_l๐P_n`$ $``$ $`cte.\underset{tI_l}{sup}\left|{\displaystyle \underset{\{n,nl\}}{}}{\displaystyle \frac{|\widehat{V}(nm)|^2}{(mn)(m+n+2t)}}\right|`$ (11)
$``$ $`cte.V_\alpha ^2{\displaystyle \frac{1}{|2n+l|^{min\{1+2\alpha ,2\}}}}.`$
We are left with the boundary terms in (6). We first discuss
$$\stackrel{~}{\stackrel{~}{๐}_l๐_l}\left(t=\frac{l}{2}\pm \frac{1}{4}\right).$$
$`\stackrel{~}{\stackrel{~}{๐}_l๐_l}=P_n๐_l\widehat{R}_l๐_l\widehat{R}_l`$, so
$$\stackrel{~}{๐}_l\stackrel{~}{๐}_l(\frac{l}{2}\pm \frac{1}{4})๐_l\widehat{R}_l(\frac{l}{2}\pm \frac{1}{4})^2cte.V_0^2\frac{1}{|2n+l|^2}$$
(12)
where we used the estimate which led to (7).
For the other boundary term we have to be more careful: Consider for $`t_{l+1}:=\frac{l}{2}+\frac{1}{4}`$
$`\sqrt{2\pi }\left(\stackrel{~}{๐}_l(t_{l+1})\stackrel{~}{๐}_{l+1}(t_{l+1})\right)(i,j)=`$
$`P_n๐(HE_n)^1(t_{l+1})(P_{n(l+1)}P_{nl})(i,j)=`$
$`\delta _{in}\delta _{j(nl)}{\displaystyle \frac{\widehat{V}(2n+l)}{3/2(2n+l+1)}}\delta _{j(n(l+1))}{\displaystyle \frac{\widehat{V}(2n+l+1)}{1/2(2n+l)}}.`$
So
$$\stackrel{~}{๐}_l(t_{l+1})\stackrel{~}{๐}_{l+1}(t_{l+1})cte.\left(\frac{|\widehat{V}(2n+l)|}{|2n+l|}+\frac{|\widehat{V}(2n+l+1)|}{|2n+l+1|}\right).$$
(13)
Furthermore it holds for $`x[\frac{1}{4},\frac{1}{4})`$:
$$\sqrt{2\pi }\stackrel{~}{๐}_l\left(\frac{l}{2}+x\right)(i,j)=\delta _{in}\frac{\widehat{V}(nj)}{(jn)(j+n+l+2x)}\chi (jn)\chi (jnl),$$
so for $`\alpha >0`$
$`\stackrel{~}{๐}_l({\displaystyle \frac{l}{2}}+x)^2`$ $``$ $`cte.{\displaystyle \underset{\{n,nl\}}{}}{\displaystyle \frac{|\widehat{V}(nj)|^2}{|(jn)(j+n+l+2x)|^2}}`$ (14)
$``$ $`V_0^2{\displaystyle \frac{cte.}{|2n+l|^2}}.`$
With these observations we have finished the proof of Theorem (2.1). We assemble the argument.
We have
$`{\displaystyle _0^t}P_nU_0^{}๐U_{L^2()}=`$
$`{\displaystyle _0^{\frac{1}{4}}}\mathrm{}+{\displaystyle \underset{l=1}{\overset{L1}{}}}{\displaystyle _{I_l}}\mathrm{}+{\displaystyle _{\frac{L}{2}\frac{1}{4}}^t}\mathrm{}`$
$`{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle _{I_l}}\mathrm{}`$
$`{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle _{I_l}}\mathrm{}_{n,l}+{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle _{I_l}}\mathrm{}_{n,l}^{}`$
where $`L`$ is the integer such that $`t[\frac{L}{2}\frac{1}{4},\frac{L}{2}+\frac{1}{4})`$; $`I_0:=[0,\frac{1}{4})`$; for the index $`L`$ we have redefined $`I_L:=[\frac{L}{2}\frac{1}{4},t)`$.
Remark that for $`n>0,\beta >1`$
$$\underset{l=0}{\overset{\mathrm{}}{}}\frac{1}{2n+l^\beta }\frac{cte.}{n^{\beta 1}}.$$
So by Lemma (3.2) it holds
$`{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle _{I_l}}\mathrm{}_{n,l}cte.\sqrt{1+V_0}{\displaystyle \frac{|\widehat{V}(2n+l)|}{\sqrt{2n+l}}}`$
$`cte.\sqrt{1+V_0}V_\alpha \left({\displaystyle \frac{1}{2n+l^{2\alpha +1}}}\right)^{1/2}`$
$`cte.V_\alpha {\displaystyle \frac{1}{n^\alpha }}.`$
By Lemma (3.3) and the estimates (7, 8, 9, 10, 11, 12, 13, 14) we have
$`{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle _{I_l}}\mathrm{}_{n,l}^{}`$
$`{\displaystyle \underset{I_l}{sup}\left(\stackrel{~}{\stackrel{~}{๐}_l๐}_l๐+\dot{\stackrel{~}{\stackrel{~}{๐}_l๐}_l}+\stackrel{~}{๐}_l๐_{n,l}+\dot{\stackrel{~}{๐}}_l\right)}+`$
$`{\displaystyle \left(\stackrel{~}{\stackrel{~}{๐}_l๐}_l(t_{l+1})+\stackrel{~}{\stackrel{~}{๐}_l๐}_l(t_l)\right)}+`$
$`{\displaystyle \left(\stackrel{~}{๐}_l(t_{l+1})\stackrel{~}{๐}_{l+1}(t_{l+1})\right)}+\stackrel{~}{๐}_0(t_0)+\stackrel{~}{๐}_L(t_L)`$
$`c_{\alpha ,M}V_\alpha {\displaystyle \frac{1}{n^{min\{1,\alpha \}}}}.`$
This proves the case $`sign(nt)>0`$. From our calculations it is clear that for $`sign(nt)<0`$ all the estimates give a bound proportional to $`V_\alpha `$ and no decay in $`n`$. Thus the proof of Theorem (2.1) is finished. $`\mathrm{}`$
## 4 Acknowledgements
This paper was partly written during the visits of G. Nenciu at CPT Marseille and of F. Bentosela and P.Duclos at Univ. of Bucharest. Financial support of CPT Marseille and the Romanian Ministry of Education (grant CNCSU 13C) is hereby acknowledged.
|
warning/0006/hep-ex0006020.html
|
ar5iv
|
text
|
# EVENT SHAPES AND POWER CORRECTIONS IN e+e- ANNIHILATION
## 1 Motivation
When $`\alpha _S`$ is determined from event shapes in e<sup>+</sup>e<sup>-</sup> annihilation the effects due to hadronisation need to be corrected. It yields a contribution to the overall error of $`\alpha _S`$ which is typically as large as the experimental systematics and the uncertainties associated with the choice of the scale. The error contribution might be alleviated by employing power corrections instead of phenomenological hadronisation models which need adjusting many parameters.
## 2 Power Corrections to Mean Values
Hadronisation is expected to cause corrections to measured observables which are suppressed by reciprocal powers of the energy scale of the process. In $`^\mathrm{?}`$ power corrections to the mean values of event shape observables are additive terms
$$=_{\mathrm{pert}}+_{\mathrm{pow}}.$$
(1)
The correction $`_{\mathrm{pow}}a_{}\alpha _0/\sqrt{s}`$ depends on a calculable observable-specific parameter $`a_{}`$ and a single non-perturbative parameter $`\alpha _0`$ to be measured experimentally which is the mean of the strong coupling $`\alpha _S`$ between 0 and 2 GeV. This type of correction has been thoroughly investigated for the thrust ($`T`$), the heavy jet mass ($`M_H`$), the total ($`B_T`$) and wide jet broadening ($`B_W`$), and the C-parameter ($`C`$) observables for $`\sqrt{s}=`$14-202 GeV. The results,$`^{\mathrm{?},\mathrm{?},\mathrm{?}}`$ updated for the corrected Milan factor,$`^\mathrm{?}`$ yield on average $`\alpha _0(2\mathrm{GeV})=0.49\pm 0.03`$ (r.m.s. 0.07) supporting the universality of $`\alpha _0`$. The r.m.s., which is larger than the combined statistical, systematic and scale uncertainties, is partly due to the neglect of the $`20\%`$ uncertainty of the Milan factor. The average of $`\alpha _S(M_\mathrm{Z})`$ is $`0.116\pm 0.004`$ which agrees with the world average.$`^\mathrm{?}`$
### 2.1 Power Corrections to $`y_3`$
Many observables are subject to power corrections of the type $`1/\sqrt{s}`$ or $`1/\sqrt{\mathrm{ln}s}`$. One observable which is known to have a leading $`1/s`$ or $`\mathrm{ln}s/s`$ correction is the 2-3-jet flip, $`y_3`$, for the $`k_{}`$ jet finder but the coefficient $`a_y`$ which determines the size of this correction is not known. Figure 1 shows the fit results of power corrections of the type $`1/\sqrt{s}`$, $`\mathrm{ln}s/\sqrt{s}`$, $`1/s`$, and $`\mathrm{ln}s/s`$ to the mean of $`y_3`$. All fits give $`\chi ^2/`$d.o.f. of about 1 but the high values $`\alpha _S(M_\mathrm{Z})=0.144\pm 0.011_{\mathrm{expt}^{}\mathrm{l}}`$ disfavour the $`1/\sqrt{s}`$-type corrections. The $`1/s`$ correction yields $`a_y=0.49\pm 0.37_{\mathrm{expt}^{}\mathrm{l}}`$ compatible with zero for $`\alpha _S(M_\mathrm{Z})=0.124\pm 0.004_{\mathrm{expt}^{}\mathrm{l}}`$ and for the first moment of $`\alpha _S`$ in the range of 0 through 2 GeV, $`\alpha _1=0.26\pm 0.02_{\mathrm{expt}^{}\mathrm{l}}`$.
### 2.2 Resummed Predictions of Mean Values
Usually all investigations used the second order calculations for the mean values of the event shape observables but the matched resummed and fixed order calculations for the distributions (see sect. 4). Resummed predictions for the mean values should give a better description of the dominating contribution from the 2-jet region to the mean value. The result of matching the resummed NLLA prediction for the mean value with the fixed order calculation is exemplified in figure 1 for the C-parameter. A comparable change of the $`๐ช(\alpha _S^2)`$ result is found for $`M_H`$. The difference is about twice as large for the $`B_W`$ while it is negligible for $`1T`$ and $`B_T`$. Fitting the data, $`\alpha _0`$ turns out to be 5-10% lower ($`40\%`$ for $`B_W`$) and $`\alpha _S(M_\mathrm{Z})`$ to be 1-10% lower if using the $`\mathrm{ln}R`$-matched resummed and fixed order calculations.
## 3 Power Corrections to Higher Moments
An straight forward extension of Eq. (1) to the second moment of the event shapes yields $`^2=_{\mathrm{pert}}^2+2_{\mathrm{pert}}_{\mathrm{pow}}+๐ช(1/s).`$ For the second moment of the thrust observable, however, a $`1/(\sqrt{s})^3`$ power correction is expected in the 2-jet region.$`^\mathrm{?}`$ The investigations $`^\mathrm{?}`$ exemplified in figure 2 show that $`1T`$, $`B_T`$, and $`C`$ require a large $`1/s`$ term which is not necessary for $`\rho ^2M_H^2/s`$ and $`B_W`$.
To suppress the $`_{\mathrm{pow}}`$ term in the formula of the second moment the study of the variance has been proposed.$`^\mathrm{?}`$ With no other data available the variance of the C-parameter, $`\sigma _C^2C^2C^2=0.034\pm 0.010`$ has been calculated from the distributions measured at 91 GeV using error propagation to assess the total error. Using the second order prediction, $`\sigma _C^20.387\alpha _S+0.0435\alpha _S^2`$, and neglecting the $`1/s`$ correction the variance yields a very low value $`\alpha _S(M_\mathrm{Z})=0.09\pm 0.03`$ which could be due to the C-parameter spectrum which does not vanish at the boundary of the 3-jet phase space.$`^\mathrm{?}`$ In general, more complete predictions are required to make use of the higher order moments.
## 4 Power Corrections to Differential Distributions
Power corrections can also be applied to the differential distributions of event shapes by shifting the resummed plus $`๐ช(\alpha _S^2)`$ prediction. This has been investi-
gated by several groups using $`T`$ and $`M_H`$. With the reanalysed JADE data at 35 and 44 GeV also power corrections to the $`C`$ and jet broadening distributions became possible which showed the necessity of squeezing the predicted distributions for the latter in addition to the shift.$`^\mathrm{?}`$ The fit results from $`T`$, $`B_T`$, $`B_W`$, and $`C`$, updated for the corrected Milan factor, yield on average $`\alpha _0=0.57\pm 0.09`$ (r.m.s. 0.12) and $`\alpha _S(M_\mathrm{Z})=0.107\pm 0.006`$. The $`\chi ^2/`$d.o.f. of the fits is about unity but for $`B_W`$ which yields $`\alpha _0=0.79`$ too high and $`\alpha _S(M_\mathrm{Z})=0.097`$ too low.
These fits of the event shape distributions exclude the extreme 2-jet region. Extending the power corrections with a shape function $`^\mathrm{?}`$ a fit over the whole distribution is possible.
## 5 Conclusions
Figure 3 summarises the results of the fits of $`\alpha _0(2\mathrm{GeV})`$ and $`\alpha _S(M_\mathrm{Z})`$ from the mean values and from the distributions of event shapes in e<sup>+</sup>e<sup>-</sup> annihilation. These results agree with those from studies of event shapes in ep scattering.$`^\mathrm{?}`$ In all these investigations power corrections prove to be a useful description of the hadronisation effects and the single non-perturbative parameter $`\alpha _0(2\mathrm{GeV})`$ assumes an universal value of about $`0.5\pm 20\%`$.
## References
|
warning/0006/gr-qc0006035.html
|
ar5iv
|
text
|
# The Bianchi IX attractor
## 1. Introduction
The last few decades, the Bianchi IX spacetimes have received considerable attention, see for instance , , and references therein. Agreement has been reached, at least concerning some aspects of the asymptotic behaviour as one approaches a singularity, but the basis for the consensus has mainly consisted of numerical studies and heuristic arguments. The objective of this article is to provide mathematical proofs for some aspects of the โacceptedโ picture. The main result of this paper was for example conjectured in p. 146-147, partly on the basis of a numerical analysis.
Why Bianchi IX? One reason is the fact that this class contains the Taub-NUT spacetimes. These spacetimes are vacuum maximal globally hyperbolic spacetimes that are causally geodesically incomplete both to the future and to the past, see and . However, as one approaches a singularity, in the sense of causal geodesic incompleteness, the curvature remains bounded. In fact, one can extend the spacetime beyond the singularities in inequivalent ways, see . It is natural to conjecture that the behaviour exhibited by the Taub-NUT spacetimes is non-generic, and it is interesting to try to prove that the behaviour is non-generic in the Bianchi IX class. In fact we prove that all Bianchi IX initial data considered in this paper other than Taub-NUT yield inextendible globally hyperbolic developments such that the curvature becomes unbounded as one approaches a singularity. This result is in fact more of an observation, since the corresponding result is known in the vacuum case, see , and curvature blow up is easy to prove in the non-vacuum cases we consider.
Another reason for studying the Bianchi IX spacetimes is the BKL conjecture, see . According to this conjecture, the โlocalโ approach to the singularity of a general solution should exhibit oscillatory behaviour. The prototypes for this behaviour among the spatially homogeneous spacetimes are the Bianchi VIII and IX classes. Furthermore the matter is conjectured to become unimportant as one approaches a singularity, with some exceptions, for example the stiff fluid case. We refer to for arguments supporting the BKL conjecture and to for an overview of conjectures and results under symmetry assumptions of varying degree. In this paper we prove, under certain restrictions on the allowed matter models, that generic Bianchi IX solutions exhibit oscillatory behaviour and that the matter becomes unimportant as one approaches a singularity. What is meant by the latter statement will be made precise below. If the matter model is a stiff fluid the matter will be important, and in that case we prove that the behaviour is quiescent. This should be compared with concerning the structure of singularities of analytic solutions to Einsteinโs equations coupled to a scalar field or stiff fluid. In that paper, Andersson and Rendall prove that given a certain kind of solution to the so called velocity dominated system, there is a unique solution of Einsteinโs equations coupled to a stiff fluid approaching the velocity dominated solution asymptotically. One can then ask the question whether it is natural to assume that a solution has the asymptotics they prescribe. In Section 20, we show that all Bianchi VIII and IX stiff fluid solutions exhibit such asymptotic behaviour.
The results presented in this paper can be divided into two parts. The first part consists of statements about developments of orthogonal perfect fluid data of class A. We clarify below what we mean by this. The results concern curvature blow up and inextendibility of developments. The second part consists of results expressed in terms of the variables of Wainwright and Hsu. These variables describe the spacetime close to the singularity, and we prove that Bianchi IX solutions generically converge to a set on which the flow of the equation coincides with the Kasner map.
We consider spatially homogeneous Lorentz manifolds $`(\overline{M},\overline{g})`$ with a perfect fluid source. The stress energy tensor is thus given by
(1.1)
$$T_{ab}=\mu u_au_b+p(\overline{g}_{ab}+u_au_b),$$
where $`u`$ is a unit timelike vectorfield, the 4-velocity of the fluid. We assume that $`p`$ and $`\mu `$ satisfy a linear equation of state
(1.2)
$$p=(\gamma 1)\mu ,$$
where we in this paper restrict our attention to $`2/3<\gamma 2`$. We will also assume that $`u`$ is perpendicular to the hypersurfaces of homogeneity. Einsteinโs equations can be written
(1.3)
$$\overline{R}_{ab}\frac{1}{2}\overline{R}\overline{g}_{ab}=T_{ab},$$
where $`\overline{R}_{ab}`$ and $`\overline{R}`$ are the Ricci and scalar curvature of $`(\overline{M},\overline{g})`$. In order to formulate an initial value problem in this setting, consider a spacelike submanifold $`(M,g)`$ of $`(\overline{M},\overline{g})`$, orthogonal to $`u`$. Let $`e_\alpha `$, $`\alpha =0,..,3`$ be a local frame with $`e_0=u`$ and $`e_i`$, $`i=1,2,3`$ tangent to $`M`$ and let $`k_{ij}`$ be the second fundamental form of $`(M,g)`$. Then $`g`$ and $`k`$ must satisfy the equations
$$R_gk_{ij}k^{ij}+(\mathrm{tr}_gk)^2=2\overline{R}_{00}+\overline{R}$$
and
$$_i\mathrm{tr}_gk^jk_{ij}=\overline{R}_{0i},$$
where $``$ is the Levi-Civita connection of $`g`$, and $`R_g`$ is the corresponding scalar curvature, indices are raised and lowered by $`g`$. If we specify a Riemannian metric $`g`$, and a symmetric covariant 2-tensor $`k`$, as initial data on a 3-manifold, they should thus in our situation satisfy
(1.4)
$$R_gk_{ij}k^{ij}+(\mathrm{tr}_gk)^2=2\mu $$
and
(1.5)
$$_i\mathrm{tr}_gk^jk_{ij}=0,$$
because of (1.3), (1.1) and the fact that $`u`$ is perpendicular to $`M`$. In other words, we should also specify the initial value of $`\mu `$ as part of the data.
We consider only a restricted class of manifolds $`M`$ and initial data. The 3-manifold $`M`$ is assumed to be a special type of Lie group, and $`g,k`$ and $`\mu `$ are assumed to be left invariant. In order to be more precise concerning the type of Lie groups $`M=G`$ we consider, let $`e_i`$, $`i=1,2,3`$ be a basis of the Lie algebra with structure constants determined by $`[e_i,e_j]=\gamma _{ij}^ke_k`$. If $`\gamma _{ik}^k=0`$, then the Lie algebra and Lie group are said to be of class A, and
(1.6)
$$\gamma _{ij}^k=ฯต_{ijm}n^{km}$$
where the symmetric matrix $`n^{ij}`$ is given by
(1.7)
$$n^{ij}=\frac{1}{2}\gamma _{kl}^{(i}ฯต^{j)kl}.$$
###### Definition 1.1.
Orthogonal perfect fluid data of class A for Einsteinโs equations consist of the following. A Lie group $`G`$ of class A, a left invariant Riemannian metric $`g`$ on $`G`$, a left invariant symmetric covariant 2-tensor $`k`$ on $`G`$, and a constant $`\mu _00`$ satisfying (1.4) and (1.5) with $`\mu `$ replaced by $`\mu _0`$.
We can choose a left invariant orthonormal basis $`\{e_i\}`$ with respect to $`g`$, so that the corresponding matrix $`n^{ij}`$ defined in (1.7) is diagonal with diagonal elements $`n_1`$, $`n_2`$ and $`n_3`$. By an appropriate choice of orthonormal basis, $`n_1,n_2,n_3`$ can be assumed to belong to one and only one of the types given in Table 1. We assign a Bianchi type to the initial data accordingly. This division constitutes a classification of the class A Lie algebras. We refer to Lemma 21.1 for a proof of these statements.
Let $`k_{ij}=k(e_i,e_j)`$. Then the matrices $`n^{ij}`$ and $`k_{ij}`$ commute according to (1.5), so that we may assume $`k_{ij}`$ to be diagonal with diagonal elements $`k_1`$, $`k_2`$ and $`k_3`$, cf. (21.13).
###### Definition 1.2.
Orthogonal perfect fluid data of class A satisfying $`k_2=k_3`$ and $`n_2=n_3`$ or one of the permuted conditions are said to be of Taub type. Data with $`\mu _0=0`$ are called vacuum data.
Observe that the Taub condition is independent of the choice of orthonormal basis diagonalizing $`n`$ and $`k`$, cf. (21.13). Considering the equations of Ellis and MacCallum (21.4)-(21.8), one can see that if $`n_2=n_3`$ and $`k_2=k_3`$ at one point in time, then the equalities always hold, cf. the construction of the spacetime carried out in the appendix. According to , vacuum solutions satisfying these conditions are the Taub-NUT solutions. This justifies the following definition.
###### Definition 1.3.
Taub-NUT initial data are type IX Taub vacuum initial data.
###### Definition 1.4.
By an orthogonal perfect fluid development of orthogonal perfect fluid data of class A, we will mean the following. A connected 4-dimensional Lorentz manifold $`(\overline{M},\overline{g})`$ and a 2-tensor $`T`$, as in (1.1), on $`(\overline{M},\overline{g})`$, such that there is an embedding $`i:G\overline{M}`$ with $`i^{}(\overline{g})=g`$, $`i^{}(\overline{k})=k`$ and $`i^{}(\mu )=\mu _0`$, where $`\overline{k}`$ is the second fundamental form of $`i(G)`$ in $`(\overline{M},\overline{g})`$.
In the appendix, we construct globally hyperbolic orthogonal perfect fluid developments, given initial data, and we refer to them as class A developments, cf. Definition 21.1. We also assign a type to such a development according to the type of the initial data. Let us make a division of the initial data according to their global behaviour.
###### Theorem 1.1.
Consider a class A development with $`1\gamma 2`$.
1. If the initial data are not of type IX, but satisfy $`\mathrm{tr}_gk=0`$, then $`\mu _0=0`$ and the development is causally geodesically complete. Only types I and VI$`I_0`$ permit this possibility.
2. If the initial data are of type I, II, V$`I_0`$, VI$`I_0`$ or VIII, and satisfy $`\mathrm{tr}_gk<0`$, then the development is future causally geodesically complete and past causally geodesically incomplete. Such initial data we will refer to as expanding.
3. Bianchi IX initial data yield developments that are past and future causally geodesically incomplete. Such data are called recollapsing.
A proof is to be found in the appendix, but observe that this theorem is not new. As far as class A developments are concerned, we will restrict our attention to equations of state with $`1\gamma 2`$. The reason is that there is cause to doubt the well posedness of the initial value problem for $`2/3<\gamma <1`$, cf. p. 85 and p. 88. Furthermore, in the Bianchi IX case we use results from concerning recollapse, see Lemma 21.6. In order to be allowed to do that, we need the above mentioned condition on $`\gamma `$. What is meant by inextendibility is explained in the following.
###### Definition 1.5.
Consider a connected Lorentz manifold $`(M,g)`$. If there is a connected $`C^2`$ Lorentz manifold $`(\widehat{M},\widehat{g})`$ of the same dimension, and a map $`i:M\widehat{M}`$, with $`i(M)\widehat{M}`$, which is an isometry onto its image, then $`(M,g)`$ is said to be $`C^2`$-extendible and $`(\widehat{M},\widehat{g})`$ is called a $`C^2`$-extension of $`(M,g)`$. A Lorentz manifold which is not $`C^2`$-extendible is said to be $`C^2`$-inextendible.
Remark. There is an analogous definition of smooth extensions. Unless otherwise mentioned, manifolds are assumed to be smooth, and maps between manifolds are assumed to be as regular as possible.
We will use the Kretschmann scalar,
(1.8)
$$\kappa =\overline{R}_{\alpha \beta \gamma \delta }\overline{R}^{\alpha \beta \gamma \delta },$$
as our main measure of whether curvature blows up or not, but in the non-vacuum case it is natural to consider the Ricci tensor contracted with itself $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$. The next theorem states the main conclusion concerning developments.
###### Theorem 1.2.
For class A developments with $`1\gamma 2`$, we have the following division.
1. Consider expanding initial data of type I, II or VI$`I_0`$ with $`1\gamma <2`$ which are not of Taub vacuum type. Then the Kretschmann scalar is unbounded along all inextendible causal geodesics in the incomplete direction.
2. Consider non-Taub-NUT recollapsing initial data with $`1\gamma <2`$. Then the Kretschmann scalar is unbounded along all inextendible causal geodesics in both incomplete directions.
3. Expanding and recollapsing data with $`\gamma =2`$ and $`\mu _0>0`$. Then the Kretschmann scalar is unbounded along all inextendible causal geodesics in all incomplete directions.
4. Expanding and recollapsing data with $`\mu _0>0`$. Then $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$ is unbounded along all inextendible causal geodesics in all incomplete directions.
In all cases mentioned above the class A development is $`C^2`$-inextendible.
Remark. Observe that the Bianchi VIII vacuum case was handled in , and the Bianchi V$`\mathrm{I}_0`$ vacuum case in . The above theorem thus isolates the vacuum Taub type solutions as the only ones among the Bianchi class A spacetimes that do not exhibit curvature blow up, given our particular matter model.
We now turn to the results that are expressed in terms of the variables of Wainwright and Hsu. The equations and some of their properties are to be found in Section 2. The appendix contains a derivation. It is natural to divide the matter models into two categories; the non-stiff fluid case and the stiff fluid case ($`\gamma =2`$).
Let us begin with the non-stiff fluid case, including the vacuum case. We confine our attention to Bianchi IX solutions. The existence interval stretches back to $`\mathrm{}`$ which corresponds to the singularity. There are some fixed points to which certain solutions converge, and data which lead to such solutions together with data of Taub type will be considered to be non-generic. The Kasner map, which is supposed to be an approximation of the Bianchi IX dynamics as one approaches a singularity, is illustrated in Figure 1. The circle in the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane appearing in the figure is called the Kasner circle, and we have depicted two bounces of the Kasner map. The starting point is marked by a star, and the end point by a plus sign. Given a point $`x`$ on the Kasner circle, the Kasner map yields a new point $`y`$ on the Kasner circle by taking the corner of the triangle closest to $`x`$, drawing a straight line from the corner through $`x`$, and then letting $`y`$ be the second point of intersection between the line and the Kasner circle. One solid line corresponds to the closure of a vacuum type II orbit of the equations of Wainwright and Hsu. Actually, it is the projection of the closure of such an orbit to the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane. A vacuum type II solution has one $`N_i`$ non-zero and the other zero, and the three different $`N_i`$ correspond to the three corners of the triangle; the rightmost corner corresponds to $`N_10`$ and the corner on the top left corresponds to $`N_30`$. The constraint (2.3) for the vacuum type II solutions is given by
$$\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\frac{3}{4}N_i^2=1.$$
The closure of this set is given a name in the following definition.
###### Definition 1.6.
The set
$$๐=\{(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},N_1,N_2,N_3):\mathrm{\Omega }+|N_1N_2|+|N_2N_3|+|N_3N_1|=0\}M,$$
where $`M`$ is defined by (2.3), is called the Bianchi attractor.
The main result of this paper is that for generic Bianchi IX data, the solution converges to the attractor. That is
(1.9)
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Omega }+N_1N_2+N_2N_3+N_3N_1)=0.$$
This conclusion supports the statement that the Kasner map approximates the dynamics, and also the statement that the matter content loses significance close to the singularity. Let us introduce some terminology.
###### Definition 1.7.
Let $`fC^{\mathrm{}}(^n,^n)`$, and consider a solution $`x`$ to the equation
$$\frac{dx}{dt}=fx,x(0)=x_0,$$
with maximal existence interval $`(t_{},t_+)`$. We call a point $`x_{}`$ an $`\alpha `$-limit point of the solution $`x`$, if there is a sequence $`t_kt_{}`$ with $`x(t_k)x_{}`$. The $`\alpha `$-limit set of $`x`$ is the set of its $`\alpha `$-limit points. The $`\omega `$-limit set is defined similarly by replacing $`t_{}`$ with $`t_+`$.
Remark. If $`t_{}>\mathrm{}`$ then the $`\alpha `$-limit set is empty, cf. .
Thus, the $`\alpha `$-limit set of a generic solution is contained in the attractor. The desired statement is that the $`\alpha `$-limit set coincides with the attractor, but the best result we have achieved in this direction is that there must at least be three $`\alpha `$-limit points on the Kasner circle. This worst case situation corresponds to the solution converging to a periodic orbit of the Kasner map with period three. Observe that we have not proven anything concerning Bianchi VIII solutions.
Let us sketch the proof. It is natural to divide it into two parts. The first part consists of proving the existence of an $`\alpha `$-limit point on the Kasner circle. We achieve this in the following steps. First we analyze the $`\alpha `$-limit sets of the Bianchi types I, II and VI$`\mathrm{I}_0`$. An analysis of types I of II can also be found in Ellis and Wainwright . Then we prove the existence of an $`\alpha `$-limit point for a generic Bianchi IX solution. To go from the existence of an $`\alpha `$-limit point to an $`\alpha `$-limit point on the Kasner circle, we use the analysis of the lower Bianchi types. In the second part, we prove (1.9). Let $`d`$ be the function appearing in that equation. We assume that $`d`$ does not converge to zero in order to reach a contradiction. The existence of an $`\alpha `$-limit point on the Kasner circle proves that there is a sequence $`\tau _k\mathrm{}`$ such that $`d(\tau _k)0`$. If $`d`$ does not converge to zero there is a $`\delta >0`$, and a sequence $`s_k\mathrm{}`$ such that $`d(s_k)\delta `$. We can assume $`s_k\tau _k`$ and conclude that $`d`$ on the whole has to grow (going backwards) in the interval $`[s_k,\tau _k]`$. What can be said about this growth? In Section 14, we prove that we can control the density parameter $`\mathrm{\Omega }`$ in this process, assuming $`\delta `$ is small enough, which is not a restriction. As a consequence $`\mathrm{\Omega }`$ can be assumed to be arbitrarily small during the growth. Some further arguments, given in Section 15, show that we can assume the growth to occur in the product $`N_2N_3`$, using the symmetries of the equations. Furthermore, one can assume the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-variables to be arbitrarily close to $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(1,0)`$, and that some expressions dominate others. For instance $`1+\mathrm{\Sigma }_+`$ can be assumed to be arbitrarily much smaller than $`N_2N_3`$. This control introduces a natural concept of order of magnitude. The behaviour of the product $`N_2N_3`$ will be oscillatory; it will look roughly like a sine wave. The point is to prove that the product decays during a period of its oscillation; that would lead to a contradiction. The variation during a period can be expressed in terms of an integral, and we use the order of magnitude concept to prove an estimate showing that this integral has the right sign.
Now consider the stiff fluid case with positive density parameter. In this case we will consider Bianchi VIII and IX solutions. The analysis is similar for the other cases and a description of the results is to be found in Section 19. Again the singularity corresponds to $`\mathrm{}`$. The density parameter $`\mathrm{\Omega }`$ converges to a non-zero value, all the $`N_i`$ converge to zero, and in the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane the solution converges to a point inside the triangle shown in Figure 2.
In Section 2, we formulate the equations of Wainwright and Hsu and briefly describe their origin and some of their properties. Section 3 contains some elementary properties of solutions. We give the existence intervals of solutions to the equations, and prove that the $`\mathrm{\Omega }\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-variables are contained in a compact set to the past for Bianchi IX solutions. As in the vacuum case, we also prove that $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ can converge to $`(1,0)`$ only if the solution is of Taub type, although this is no longer a characterization. In Section 4, we mention some critical points and make more precise the statement that solutions converging to these points are non-generic. Included in this section are also two technical lemmas relevant to the analysis. The monotonicity principle is explained in Section 5. It is fundamental to the analysis of the $`\alpha `$-limit sets of the solutions. We present two applications; the fact that all $`\alpha `$-limit points of Bianchi IX solutions are of type I, II or VI$`\mathrm{I}_0`$ and an analysis of the vacuum type II orbits. The last application is not complicated, but illustrates the arguments involved as well as demonstrating how the map depicted in Figure 1 can be viewed as a sequence of type II orbits. Section 6 deals with situations such that one has control over the shear variables and the density parameter. Specifically, it gives a geometric interpretation of some of the equations in $`\mathrm{\Omega }\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-space. As an application, we prove that if a Bianchi IX solution has an $`\alpha `$-limit point on the Kasner circle then all the points obtained by applying the Kasner map to this point belong to the $`\alpha `$-limit set of the solution. The stiff fluid case is handled in Section 7. In this case the $`\alpha `$-limit set consists of a point regardless of type. Sections 8-10 deal with the lower order Bianchi types needed in order to analyze Bianchi IX. An analysis of types I of II can also be found in Ellis and Wainwright . Section 11 gives the possibilities for a Taub type Bianchi IX solution. The technical Section 12 is needed in order to prove the existence of an $`\alpha `$-limit point for Bianchi IX solutions, and also to prove that the set of vacuum type II points is an attractor. It is used for approximating the solution in situations where the behaviour is oscillatory. Section 13 proves the existence of an $`\alpha `$-limit point for a Bianchi IX solution and the existence of an $`\alpha `$-limit point on the Kasner circle for generic Bianchi IX solutions. In Section 14, we prove that if one has control over the sum $`|N_1N_2|+|N_2N_3|+|N_3N_1|`$ in some time interval $`[\tau _1,\tau _2]`$, and control over $`\mathrm{\Omega }`$ in $`\tau _2`$ then one has control over $`\mathrm{\Omega }`$ in the entire interval. This rather technical observation is essential in the proof that generic solutions converge to the attractor. The heart of this paper is Section 15 which contains a proof of (1.9). It also contains arguments that will be used in Section 16 to analyze the regularity of the set of non-generic points. In Section 17, we observe that the convergence to the attractor is uniform, and in Section 18 we prove the existence of at least three non-special $`\alpha `$-limit points on the Kasner circle. We formulate the main conclusions and prove Theorem 1.2 in Section 19. In Section 20, we relate our results concerning stiff fluid solutions to those of . The appendices contain results relating solutions to the equations of Wainwright and Hsu with properties of the class A developments and some curvature computations.
## 2. Equations of Wainwright and Hsu
The essence of this paper is an analysis of the asymptotic behaviour of solutions to the equations of Wainwright and Hsu (2.1)-(2.3). One important property of these equations is that they describe all the Bianchi class A types at the same time. Another important property is that it seems that the variables remain in a compact set as one approaches a singularity. In the Bianchi IX case, this follows from the analysis presented in this paper. Let us give a rough description of the origin of the variables. In the situations we consider, there is a foliation of the Lorentz manifold by homogeneous spacelike hypersurfaces diffeomorphic to a Lie group $`G`$ of class A. One can define an orthonormal basis $`e_\alpha `$, $`\alpha =0,\mathrm{},3`$, such that $`e_i`$, $`i=1,2,3`$, span the tangent space of the spacelike hypersurfaces of homogeneity, and $`e_0=_t`$ for a suitable globally defined time coordinate $`t`$. It is possible to associate a matrix $`n_{ij}`$ with the spacelike vectors $`e_i`$, as in (1.7), and assume it to be diagonal with diagonal components $`n_i`$. One changes the time coordinate by $`dt/d\tau =3/\theta `$, where $`\theta `$ is minus the trace of the second fundamental form of the spacelike hypersurface corresponding to $`t`$. The $`N_i(\tau )`$ below are the $`n_i(\tau )`$ divided by $`\theta (\tau )`$, the $`\mathrm{\Sigma }_+`$ and $`\mathrm{\Sigma }_{}`$ correspond to the traceless part of the second fundamental form of the spacelike hypersurface corresponding to $`\tau `$, similarly normalized, and finally $`\mathrm{\Omega }=3\mu /\theta ^2`$. We will refer to $`\mathrm{\Sigma }_+`$ and $`\mathrm{\Sigma }_{}`$ as the shear variables, and to $`\mathrm{\Omega }`$ as the density parameter. The question then arises to what extent this makes sense, since $`\theta `$ could become zero. An answer is given in the appendix. For all the Bianchi types except IX, this procedure is essentially harmless, and the variables of Wainwright and Hsu capture the entire Lorentz manifold. In the Bianchi IX case, there is however a point at which $`\theta =0`$, at least if $`1\gamma 2`$, see the appendix, and the variables are only valid for half a development in that case. As far as the analysis of the asymptotics are concerned, this is however not important. A derivation of the equations is given in the appendix. They are
$`N_1^{}`$ $`=`$ $`(q4\mathrm{\Sigma }_+)N_1`$
$`N_2^{}`$ $`=`$ $`(q+2\mathrm{\Sigma }_++2\sqrt{3}\mathrm{\Sigma }_{})N_2`$
(2.1) $`N_3^{}`$ $`=`$ $`(q+2\mathrm{\Sigma }_+2\sqrt{3}\mathrm{\Sigma }_{})N_3`$
$`\mathrm{\Sigma }_+^{}`$ $`=`$ $`(2q)\mathrm{\Sigma }_+3S_+`$
$`\mathrm{\Sigma }_{}^{}`$ $`=`$ $`(2q)\mathrm{\Sigma }_{}3S_{}`$
$`\mathrm{\Omega }^{}`$ $`=`$ $`[2q(3\gamma 2)]\mathrm{\Omega }.`$
The prime denotes derivative with respect to a time coordinate $`\tau `$, and
$`q`$ $`=`$ $`{\displaystyle \frac{1}{2}}(3\gamma 2)\mathrm{\Omega }+2(\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2)`$
(2.2) $`S_+`$ $`=`$ $`{\displaystyle \frac{1}{2}}[(N_2N_3)^2N_1(2N_1N_2N_3)]`$
$`S_{}`$ $`=`$ $`{\displaystyle \frac{\sqrt{3}}{2}}(N_3N_2)(N_1N_2N_3).`$
The constraint is
(2.3)
$$\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\frac{3}{4}[N_1^2+N_2^2+N_3^22(N_1N_2+N_2N_3+N_3N_1)]=1.$$
We demand that $`2/3<\gamma 2`$ and $`\mathrm{\Omega }0`$. The equations (2.1)-(2.3) have certain symmetries, described in Wainwright and Hsu . By permuting $`N_1,N_2,N_3`$ arbitrarily, we get new solutions, if we at the same time carry out appropriate combinations of rotations by integer multiples of $`2\pi /3`$, and reflections in the $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$-plane. Explicitly, the transformations
$$(\stackrel{~}{N}_1,\stackrel{~}{N}_2,\stackrel{~}{N}_3)=(N_3,N_1,N_2),(\stackrel{~}{\mathrm{\Sigma }}_+,\stackrel{~}{\mathrm{\Sigma }}_{})=(\frac{1}{2}\mathrm{\Sigma }_++\frac{1}{2}\sqrt{3}\mathrm{\Sigma }_{},\frac{1}{2}\sqrt{3}\mathrm{\Sigma }_+\frac{1}{2}\mathrm{\Sigma }_{})$$
and
$$(\stackrel{~}{N}_1,\stackrel{~}{N}_2,\stackrel{~}{N}_3)=(N_1,N_3,N_2),(\stackrel{~}{\mathrm{\Sigma }}_+,\stackrel{~}{\mathrm{\Sigma }}_{})=(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})$$
yield new solutions. Below, we refer to rotations by integer multiples of $`2\pi /3`$ as rotations. Changing the sign of all the $`N_i`$ at the same time does not change the equations. Classify points $`(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},N_1,N_2,N_3)`$ according to the values of $`N_1,N_2,N_3`$ in the same way as in Table 1. Since the sets $`N_i>0`$, $`N_i<0`$ and $`N_i=0`$ are invariant under the flow of the equations, we may classify solutions to (2.1)-(2.3) accordingly.
###### Definition 2.1.
The Kasner circle is defined by the conditions $`N_i=\mathrm{\Omega }=0`$ and the constraint (2.3). There are three points on this circle called special: $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(1,0)`$ and $`(1/2,\pm \sqrt{3}/2)`$.
The following reformulation of $`\mathrm{\Sigma }_+^{}`$ is written down for future reference,
(2.4)
$$\mathrm{\Sigma }_+^{}=(22\mathrm{\Omega }2\mathrm{\Sigma }_+^22\mathrm{\Sigma }_{}^2)(\mathrm{\Sigma }_++1)\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_++\frac{9}{2}N_1(N_1N_2N_3).$$
## 3. Elementary properties of solutions
Here we collect some miscellaneous observations that will be of importance. Most of them are similar to results obtained in . The $`\alpha `$-limit set defined in Definition 1.7 plays an important role in this paper, and here we mention some of its properties.
###### Lemma 3.1.
Let $`f`$ and $`x`$ be as in Definition 1.7. The $`\alpha `$-limit set of $`x`$ is closed and invariant under the flow of $`f`$. If there is a $`T`$ such that $`x(t)`$ is contained in a compact set for $`tT`$, then the $`\alpha `$-limit set of $`x`$ is connected.
Proof. See e. g. . $`\mathrm{}`$
###### Definition 3.1.
A solution to (2.1)-(2.3) satisfying $`N_2=N_3`$ and $`\mathrm{\Sigma }_{}=0`$, or one of the conditions found by applying the symmetries, is said to be of Taub type.
Remark. The set defined by $`N_2=N_3`$ and $`\mathrm{\Sigma }_{}=0`$ is invariant under the flow of (2.1).
###### Lemma 3.2.
The existence intervals for all solutions to (2.1)-(2.3) except Bianchi IX are $`(\mathrm{},\mathrm{})`$. For Bianchi IX solutions we have past global existence.
Proof. As in the vacuum case, see . $`\mathrm{}`$
By observations made in the appendix, $`\mathrm{}`$ corresponds to the singularity.
###### Lemma 3.3.
Let $`2/3<\gamma 2`$. Consider a solution of type IX. The image $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},\mathrm{\Omega })((\mathrm{},0])`$ is contained in a compact set whose size depends on the initial data. Further, if at a point in time $`N_3N_2N_1`$ and $`N_32`$, then $`N_2N_3/10`$.
Proof. As in the vacuum case, see . $`\mathrm{}`$
That $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},\mathrm{\Omega })`$ is contained in a compact set for all the other types follows from the constraint. The second part of this lemma will be important in the proof of the existence of an $`\alpha `$-limit point. One consequence is that one $`N_i`$ may not become unbounded alone.
The final observation is relevant in proving curvature blow up. One can define a normalized version (22.3) of the Kretschmann scalar (1.8), and it can be expressed as a polynomial in the variables of Wainwright and Hsu. One way of proving that a specific solution exhibits curvature blow up is to prove that it has an $`\alpha `$-limit point at which the normalized Kretschmann scalar is non-zero. We refer to the appendix for the details. It turns out that this polynomial is zero when $`N_2=N_3`$, $`N_1=0`$, $`\mathrm{\Sigma }_{}=0`$, $`\mathrm{\Sigma }_+=1`$ and $`\mathrm{\Omega }=0`$. The same is true of the points obtained by applying the symmetries. It is then natural to ask the question: for which solutions does $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ converge to $`(1,0)`$?
###### Proposition 3.1.
A solution to (2.1)-(2.3) with $`2/3<\gamma <2`$ satisfies
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Sigma }_+(\tau ),\mathrm{\Sigma }_{}(\tau ))=(1,0),$$
only if it is contained in the invariant set $`\mathrm{\Sigma }_{}=0`$ and $`N_2=N_3`$.
Remark. The proposition does not apply to the stiff fluid case. The analogous statements for the points $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(1/2,\pm \sqrt{3}/2)`$ are true by an application of the symmetries. We may not replace the implication with an equivalence, cf. Proposition 9.1.
Proof. The argument is essentially the same as in the vacuum case, see . We only need to observe that $`\mathrm{\Omega }`$ will decay exponentially when $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ is close to $`(1,0)`$. $`\mathrm{}`$
## 4. Critical points
###### Definition 4.1.
The critical point $`F`$ is defined by $`\mathrm{\Omega }=1`$ and all other variables zero. In the case $`2/3<\gamma <2`$, we define the critical point $`P_1^+(II)`$ to be the type II point with $`\mathrm{\Sigma }_{}=0`$, $`N_1>0`$, $`\mathrm{\Sigma }_+=(3\gamma 2)/8`$ and $`\mathrm{\Omega }=1(3\gamma 2)/16`$. The critical points $`P_i^+(II)`$, $`i=2,3`$ are found by applying the symmetries.
It will turn out that there are solutions which converge to these points as $`\tau \mathrm{}`$. The main objective of this section is to prove that the set of such solutions is small. Observe that only non-vacuum solutions can converge these critical points.
###### Definition 4.2.
Let $`_{\mathrm{VII}_0}`$ denote initial data to (2.1)-(2.3) of type VI$`\mathrm{I}_0`$ with $`\mathrm{\Omega }>0`$, and correspondingly for the other types. Let $`๐ซ_{\mathrm{VII}_0}`$ be the elements of $`_{\mathrm{VII}_0}`$ such that the corresponding solutions converge to one of $`P_i^+(II)`$ as $`\tau \mathrm{}`$ and similarly for Bianchi II and IX. Finally, let $`_{\mathrm{VII}_0}`$ be the elements of $`_{\mathrm{VII}_0}`$ such that the corresponding solutions converge to $`F`$ as $`\tau \mathrm{}`$, and similarly for the other types.
Remark. The sets $`_{\mathrm{II}}`$ and so on depend on $`\gamma `$, but we omit this reference.
Observe that $`_\mathrm{I}`$, $`_{\mathrm{II}}`$, $`_{\mathrm{VII}_0}`$ and $`_{\mathrm{IX}}`$ are submanifolds of $`^6`$ of dimensions 2, 3, 4 and 5 respectively. They are diffeomorphic with open sets in a suitable $`^n`$; project $`\mathrm{\Omega }`$ to zero. We will prove that $`๐ซ_{\mathrm{II}}`$ consists of points and that $`_\mathrm{I}`$ is the point $`F`$. Let $`2/3<\gamma <2`$ be fixed. In Theorem 16.1, we will be able to prove that the sets $`_{\mathrm{II}},_{\mathrm{VII}_0}`$, $`_{\mathrm{IX}}`$, $`๐ซ_{\mathrm{VII}_0}`$ and $`๐ซ_{\mathrm{IX}}`$ are $`C^1`$ submanifolds of $`^6`$ of dimensions $`1`$, $`2`$, $`3`$, $`1`$ and $`2`$ respectively. This justifies the following definition.
###### Definition 4.3.
Let $`2/3<\gamma <2`$. A solution to (2.1)-(2.3) is said to be generic if it is not of Taub type, and if it does not belong to $`_\mathrm{I},_{\mathrm{II}},_{\mathrm{VII}_0}`$, $`_{\mathrm{IX}}`$, $`๐ซ_{\mathrm{II}}`$, $`๐ซ_{\mathrm{VII}_0}`$ or $`๐ซ_{\mathrm{IX}}`$.
We will need the following two lemmas in the sequel.
###### Lemma 4.1.
Consider a solution $`x`$ to (2.1)-(2.3) such that $`x`$ has $`P_1^+(II)`$ as an $`\alpha `$-limit point but does not converge to it. Then $`x`$ has an $`\alpha `$-limit point of type II, which is not $`P_1^+(II)`$.
Remark. There is no solution satisfying the conditions of this lemma, but we will need it to establish that fact.
Proof. Consider the solution to belong to $`^6`$, and let the point $`x_0`$ represent $`P_1^+(II)`$. There is an $`ฯต>0`$ such that for each $`T`$, there is a $`\tau T`$ such that $`x(\tau )`$ does not belong to the open ball $`B_ฯต(x_0)`$. In $`x_0`$ one can compute that
$$q+2\mathrm{\Sigma }_+\pm 2\sqrt{3}\mathrm{\Sigma }_{}>0.$$
Let $`ฯต`$ be so small that these expressions are positive in $`B_ฯต(x_0)`$. Let $`\tau _k\mathrm{}`$ be a sequence such that $`x(\tau _k)x_0`$, and let $`s_k\tau _k`$ be a sequence such that $`x(s_k)B_ฯต(x_0)`$ and $`x((s_k,\tau _k])B_ฯต(x_0)`$. Since $`x(s_k)`$ is contained in a compact set, there is a convergent subsequence yielding an $`\alpha `$-limit point which is not $`P_1^+(II)`$. Since $`N_2`$ and $`N_3`$ converge to zero in $`\tau _k`$ and decay in absolute value from $`\tau _k`$ to $`s_k`$, the $`\alpha `$-limit point has to be of type II ($`N_1`$ has to be non-zero for the new $`\alpha `$-limit point if $`ฯต`$ is small enough). $`\mathrm{}`$
###### Lemma 4.2.
Consider a solution $`x`$ to (2.1)-(2.3) such that $`x`$ has $`F`$ as an $`\alpha `$-limit point, but which does not converge to $`F`$. Then $`x`$ has an $`\alpha `$-limit point of type I which is not $`F`$.
Remark. The same remark as that made in connection with Lemma 4.1 holds concerning this lemma.
Proof. The idea is the same as the previous lemma. We need only observe that $`q4\mathrm{\Sigma }_+,q+2\mathrm{\Sigma }_++2\sqrt{3}\mathrm{\Sigma }_{}`$ and $`q+2\mathrm{\Sigma }_+2\sqrt{3}\mathrm{\Sigma }_{}`$ are positive in $`F`$. $`\mathrm{}`$
## 5. The monotonicity principle
The following lemma will be a basic tool in the analysis of the asymptotics, we will refer to it as the monotonicity principle.
###### Lemma 5.1.
Consider
(5.1)
$$\frac{dx}{dt}=fx$$
where $`fC^{\mathrm{}}(^n,^n)`$. Let $`U`$ be an open subset of $`^n`$, and $`M`$ a closed subset invariant under the flow of the vectorfield $`f`$. Let $`G:U`$ be a continuous function such that $`G(x(t))`$ is strictly monotone for any solution $`x(t)`$ of (5.1), as long as $`x(t)UM`$. Then no solution of (5.1) whose image is contained in $`UM`$ has an $`\alpha `$\- or $`\omega `$-limit point in $`U`$.
Remark. Observe that one can use $`M=^n`$. We will mainly choose $`M`$ to be the closed invariant subset of $`^6`$ defined by (2.3). If one $`N_i`$ is zero and two are non-zero, we consider the number of variables to be four etc.
Proof. Suppose $`pU`$ is an $`\alpha `$-limit point of a solution $`x`$ contained in $`UM`$. Then $`Gx`$ is strictly monotone. There is a sequence $`t_nt_{}`$ such that $`x(t_n)p`$ by our supposition. Thus $`G(x(t_n))G(p)`$, but $`Gx`$ is monotone so that $`G(x(t))G(p)`$. Thus $`G(q)=G(p)`$ for all $`\alpha `$-limit points $`q`$ of $`x`$. Since $`M`$ is closed $`pM`$. The solution $`\overline{x}`$ of (5.1), with initial value $`p`$, is contained in $`M`$ by the invariance property of $`M`$, and it consists of $`\alpha `$-limit points of $`x`$ so that $`G(\overline{x}(t))=G(p)`$ which is constant. Furthermore, on an open set containing zero it takes values in $`U`$ contradicting the assumptions of the lemma. $`\mathrm{}`$
Let us give an example of an application.
###### Lemma 5.2.
Consider a solution to (2.1)-(2.3) of type VIII or IX. If it has an $`\alpha `$-limit point, then
$$\underset{\tau \mathrm{}}{lim}(N_1N_2N_3)(\tau )=0.$$
Proof. Let $`U`$ of Lemma 5.1 be defined by the union of the sets $`N_i0`$, $`i=1,2,3`$, $`M`$ by the constraint (2.3), and $`G`$ by the function $`N_1N_2N_3`$. Compute
(5.2)
$$(N_1N_2N_3)^{}=3qN_1N_2N_3.$$
Consider a solution $`x`$ of (2.1)-(2.3). We need to prove that $`Gx`$ is strictly monotone as long as $`x(\tau )UM`$. By (5.2) the only problem that could occur is $`q=0`$. However, $`q=0`$ implies $`|\mathrm{\Sigma }_+^{}|+|\mathrm{\Sigma }_{}^{}|>0`$ by (2.1)-(2.3) so that $`Gx`$ has the desired property. If the sequence $`\tau _k\mathrm{}`$ yields the $`\alpha `$-limit point we assume exists, then we conclude that
$$(N_1N_2N_3)(\tau _k)0.$$
Since $`N_1N_2N_3`$ is monotone, we conclude that it converges to zero. $`\mathrm{}`$
One important consequence of this observation is the fact that all $`\alpha `$-limit points of Bianchi VIII and IX solutions are of one of the lower Bianchi types. Since the $`\alpha `$-limit set is invariant under the flow, it is thus of interest to know something about the $`\alpha `$-limit sets of the lower Bianchi types, if one wants to prove the existence of an $`\alpha `$-limit point on the Kasner circle.
Let us now analyze the vacuum type II orbits and define the Kasner map.
###### Proposition 5.1.
A Bianchi II vacuum solution of (2.1)-(2.3) with $`N_1>0`$ and $`N_2=N_3=0`$ satisfies
(5.3)
$$\underset{\tau \pm \mathrm{}}{lim}N_1=0.$$
The $`\omega `$-limit set is a point in $`๐ฆ_1`$ and the $`\alpha `$-limit set is a point on the Kasner circle, in the complement of the closure of $`๐ฆ_1`$.
Remark. What is meant by $`๐ฆ_1`$ is explained in Definition 6.1.
Proof. Using the constraint (2.3) we deduce that
$$\mathrm{\Sigma }_+^{}=\frac{3}{2}N_1^2(2\mathrm{\Sigma }_+).$$
We wish to apply the monotonicity principle. There are three variables. Let $`U`$ be defined by $`N_1>0`$, $`M`$ be defined by (2.3), and $`G(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},N_1)=\mathrm{\Sigma }_+`$. We conclude that (5.3) is true as follows. Let $`\tau _n\mathrm{}`$. A subsequence yields an $`\omega `$-limit point by (2.3). The monotonicity principle yields $`N_1(\tau _{n_k})0`$ for the subsequence. The argument for the $`\alpha `$-limit set is similar, and equation (5.3) follows. Combining this with the constraint, we deduce
$$\underset{\tau \pm \mathrm{}}{lim}q=2.$$
Using the monotonicity of $`\mathrm{\Sigma }_+`$, we conclude that $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ has to converge. As for the $`\alpha `$-limit set, convergence to $`๐ฆ_1`$ is not allowed since $`N_1^{}<0`$ close to $`๐ฆ_1`$. Convergence to one of the special points in the closure of $`๐ฆ_1`$ is also forbidden, since Proposition 3.1 would imply $`N_1=0`$ for the solution in that case. Assume now that $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})(\sigma _+,\sigma _{})`$ as $`\tau \mathrm{}`$. Compute
(5.4)
$$\left(\frac{\mathrm{\Sigma }_{}}{2\mathrm{\Sigma }_+}\right)^{}=0.$$
We get
(5.5)
$$\frac{\mathrm{\Sigma }_{}}{2\mathrm{\Sigma }_+}=\frac{\sigma _{}}{2\sigma _+}$$
for arbitrary $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ belonging to the solution. Since $`N_1^{}=(q4\mathrm{\Sigma }_+)N_1`$ and $`N_10`$, we have to have $`\sigma _+1/2`$. If $`\sigma _+=1/2`$, then $`\sigma _{}=\pm \sqrt{3}/2`$. The two corresponding lines in the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane, obtained by substituting $`(\sigma _+,\sigma _{})`$ into (5.5), do not intersect any points interior to the Kasner circle. Therefore $`\sigma _+=1/2`$ is not an allowed limit point, and the proposition follows. $`\mathrm{}`$
Observe that by (5.4), the projection of the solution to the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane is a straight line. The orbits when $`N_2>0`$ and when $`N_3>0`$ are obtained by applying the symmetries. Figure 1 shows a sequence of vacuum type II orbits projected to the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane. The first line, starting at the star, has $`N_1>0`$, the second $`N_3>0`$ and the third $`N_2>0`$.
###### Definition 5.1.
If $`x_0`$ is a non-special point on the Kasner circle, then the Kasner map applied to $`x_0`$ is defined to be the point $`x_1`$ on the Kasner circle, with the property that there is a vacuum type II orbit with $`x_0`$ as an $`\omega `$-limit point and $`x_1`$ as an $`\alpha `$-limit point.
## 6. Dependence on the shear variables
In several arguments, we will have control over the shear variables and the density parameter in some time interval, and it is of interest to know how the remaining variables behave in such situations. Consider for instance the expression multiplying $`N_1`$ in the formula for $`N_1^{}`$, see (2.1). It is given by $`q4\mathrm{\Sigma }_+`$ and equals zero when
(6.1)
$$\frac{1}{4}(3\gamma 2)\mathrm{\Omega }+(1\mathrm{\Sigma }_+)^2+\mathrm{\Sigma }_{}^2=1.$$
The set of points in $`\mathrm{\Omega }\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-space satisfying this equation is a paraboloid, and the intersection with $`\mathrm{\Omega }=0`$ is the dashed circle shown in Figure 3. If $`(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ belongs to the interior of the paraboloid (6.1) with $`\mathrm{\Omega }0`$, then $`|N_1|^{}`$ will be negative, so that $`|N_1|`$ increases as we go backward. Outside of the paraboloid, $`|N_1|`$ decreases. The situation is similar for $`N_2`$ and $`N_3`$. Observe that the circle obtained by letting $`\mathrm{\Omega }=0`$ in (6.1) intersects the Kasner circle in two special points. The same is true of the rotated circles corresponding to $`N_2`$ and $`N_3`$. It will be convenient to introduce notation for the points on the Kasner circle at which $`|N_i|^{}`$ is negative.
###### Definition 6.1.
We let $`๐ฆ_1,๐ฆ_2`$ and $`๐ฆ_3`$ be the subsets of the Kasner circle where $`q4\mathrm{\Sigma }_+<0,q+2\mathrm{\Sigma }_++2\sqrt{3}\mathrm{\Sigma }_{}<0`$ and $`q+2\mathrm{\Sigma }_+2\sqrt{3}\mathrm{\Sigma }_{}<0`$ respectively.
Remark. On the Kasner circle, $`\mathrm{\Omega }=0`$ so that $`q=2(\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2)=2`$ under the conditions of this definition.
It also of interest to know when the derivatives of $`N_2N_3`$ and similar products are zero. Since $`(N_2N_3)^{}=(2q+4\mathrm{\Sigma }_+)N_2N_3`$, we consider the set on which $`q+2\mathrm{\Sigma }_+`$ equals zero. This set is a paraboloid and is given by
$$\frac{1}{4}(3\gamma 2)\mathrm{\Omega }+(\mathrm{\Sigma }_++\frac{1}{2})^2+\mathrm{\Sigma }_{}^2=\frac{1}{4}.$$
The intersection with the plane $`\mathrm{\Omega }=0`$ is the circle with radius $`1/2`$ shown in Figure 3. Again, inside the paraboloid $`|N_2N_3|`$ increases as we go backward, and outside it decreases. There are corresponding paraboloids for the products $`N_1N_2`$ and $`N_1N_3`$. Observe that in the non-vacuum case, it is harmless to introduce $`\omega =\mathrm{\Omega }^{1/2}`$ and then the paraboloids become half spheres.
###### Proposition 6.1.
Consider a Bianchi IX solution to (2.1)-(2.3) with $`2/3<\gamma <2`$. If the solution has a non-special $`\alpha `$-limit point $`x`$ on the Kasner circle, then the closure of the vacuum type II orbit with $`x`$ as an $`\omega `$-limit point belongs to the $`\alpha `$-limit set.
Remark. The same conclusion holds for a Bianchi type VI$`\mathrm{I}_0`$ solution with $`N_1=0`$, if it has an $`\alpha `$-limit point in $`๐ฆ_2`$ or $`๐ฆ_3`$.
Proof. Assume the limit point lies in $`๐ฆ_1`$ with $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(\sigma _+,\sigma _{})`$. There is a sequence $`\tau _k\mathrm{}`$, such that the solution evaluated at $`\tau _k`$ converges to the point on the Kasner circle. There is a ball $`B_\eta (\sigma _+,\sigma _{})`$ in the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane, centered at this point, such that $`|N_2|,|N_3|,|N_1N_2|,|N_1N_3|`$ and $`\mathrm{\Omega }`$ all decay exponentially, at least as $`e^{\xi \tau }`$ for some fixed $`\xi >0`$, and $`N_1`$ increases exponentially, at least as $`e^{\xi \tau }`$, in the closure of this ball. There is a $`K`$ such that $`(\mathrm{\Sigma }_+(\tau _k),\mathrm{\Sigma }_{}(\tau _k))B_\eta (\sigma _+,\sigma _{})`$ for all $`kK`$. For each time we enter the ball, we must leave it, since if we stay in it to the past, $`N_1`$ will grow to infinity whereas $`N_2`$ and $`N_3`$ will decay to zero, in violation of the constraint. Thus for each $`\tau _k`$, $`kK`$, there is a $`t_k\tau _k`$ corresponding to the first time we leave the ball, starting at $`\tau _k`$ and going backward. We may compute
$$(\frac{\mathrm{\Sigma }_{}}{2\mathrm{\Sigma }_+})^{}=h$$
where
$$|h(\tau )|ฯต_ke^{\xi (\tau \tau _k)}$$
in $`[t_k,\tau _k]`$ and $`ฯต_k0`$. Thus
$$\frac{\mathrm{\Sigma }_{}(\tau _k)}{2\mathrm{\Sigma }_+(\tau _k)}\frac{\mathrm{\Sigma }_{}(t_k)}{2\mathrm{\Sigma }_+(t_k)}=_{t_k}^{\tau _k}h๐\tau .$$
But
$$|_{t_k}^{\tau _k}h๐\tau |\frac{ฯต_k}{\xi },$$
and in consequence
$$\frac{\mathrm{\Sigma }_{}(\tau _k)}{2\mathrm{\Sigma }_+(\tau _k)}\frac{\mathrm{\Sigma }_{}(t_k)}{2\mathrm{\Sigma }_+(t_k)}0.$$
We thus get a type II vacuum limit point with $`N_1>0`$, to which we may apply the flow, and deduce the conclusion of the lemma. The statement made in the remark follows in the same way. Observe that the only important thing was that the limit point was in $`๐ฆ_1`$ and $`N_1`$ was non-zero for the solution. $`\mathrm{}`$
## 7. The stiff fluid case
In this section we will assume $`\mathrm{\Omega }>0`$ and $`\gamma =2`$ for all solutions we consider. We begin by explaining the origin of the triangle shown in Figure 2. Then we analyze the type II orbits. They yield an analogue of the Kasner map, connecting two points inside the Kasner circle, and we state an analogue of Proposition 6.1 for this map. We then prove that $`\mathrm{\Omega }`$ is bounded away from zero to the past. Only in the case of Bianchi IX is an argument required, but this result is the central part of the analysis of the stiff fluid case. A peculiarity of the equations then yields the conclusion that $`|N_1N_2|+|N_2N_3|+|N_3N_1|`$ converges to zero exponentially. This proves that any solution is contained in a compact set to the past, and that all $`\alpha `$-limit points are of type I or II. Another consequence is that $`\mathrm{\Omega }`$ has to converge to a non-zero value; this requires a proof in the Bianchi IX case. Next one concludes that all $`N_i`$ converge to zero, since if that were not the case, there would be an $`\alpha `$-limit point of type II to which one could apply the flow, obtaining $`\alpha `$-limit points with different $`\mathrm{\Omega }`$:s. Then if a Bianchi IX solution had an $`\alpha `$-limit point outside the triangle, one could apply the โKasnerโ map to such a point, obtaining an $`\alpha `$-limit point with some $`N_i>0`$. Finally, some technical arguments finish the analysis.
In the case of a stiff fluid, that is $`\gamma =2`$, it is convenient to introduce
$$\omega =\mathrm{\Omega }^{1/2}.$$
We then have, since $`3\gamma 2=4`$,
(7.1)
$$\omega ^{}=(2q)\omega .$$
The expression $`\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2`$ turns into $`\omega ^2+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2`$, and the $`\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{}`$-coordinates of the type I points obey
(7.2)
$$\omega ^2+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2=1,\omega 0.$$
In the stiff fluid case, all the type I points are fixed points, and they play a role similar to that of the Kasner circle in the vacuum case.
Let us make some observations. If $`N_10`$, then $`N_1^{}=0`$ is equivalent to $`q4\mathrm{\Sigma }_+=0`$. Dividing by $`2`$ and completing squares, we see that this condition is equivalent to
(7.3)
$$\omega ^2+(1\mathrm{\Sigma }_+)^2+\mathrm{\Sigma }_{}^2=1,\omega 0.$$
By applying the symmetries, the conditions $`N_i^{}=0,N_i0`$ are consequently all fulfilled precisely on half spheres of radii $`1`$. Since $`|N_1|^{}<0`$ corresponds to an increase in $`|N_1|`$ as we go backward, $`|N_1|`$ increases exponentially as we are inside the half sphere (7.3) and decreases exponentially as we are outside it. If one takes the intersection of (7.2) and (7.3), one gets the subset $`\mathrm{\Sigma }_+=1/2`$ of (7.2). The corresponding intersections for $`N_2`$ and $`N_3`$ yield two more lines in the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane. Together they yield the triangle in Figure 2. Consequently, if $`(\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ is close to (7.2) and $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ is in the interior of the triangle, then all the $`N_i`$ decay exponentially as $`\tau \mathrm{}`$.
Let $`_1`$ be the subset $`\omega \mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-space obeying (7.2) with $`\omega >0`$ and $`\mathrm{\Sigma }_+>1/2`$ and $`_2`$, $`_3`$ be the corresponding sets for $`N_2`$ and $`N_3`$. We also let $`_1`$ be the subset of the intersection between (7.2) and (7.3) with $`\omega >0`$ and correspondingly $`N_2`$ and $`N_3`$ yield $`_2`$ and $`_3`$.
###### Lemma 7.1.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ such that $`N_1>0`$, $`\omega >0`$ and $`N_2=N_3=0`$. Then
(7.4)
$$\underset{\tau \pm \mathrm{}}{lim}N_1(\tau )=0$$
and $`(\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ converges to a point, satisfying (7.2) and $`\omega >0`$, in the complement of $`_1_1`$, as $`\tau \mathrm{}`$. In $`\omega \mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-space, the orbit of the solution is a straight line connecting two points satisfying (7.2). If $`\omega >0`$, it is strictly increasing along the solution, going backwards in time.
Proof. Since $`q<2`$ for the entire solution, we can apply the monotonicity principle with $`U`$ defined by $`q<2`$, $`G`$ defined by $`\mathrm{\Sigma }_+`$ and $`M`$ by the constraint (2.3). If $`q`$ does not converge to 2 as $`\tau \mathrm{}`$, we get an $`\alpha `$-limit point with $`q<2`$. We have a contradiction. This argument also yields the conclusion that $`N_10`$ as $`\tau \mathrm{}`$. Equation (7.4) follows. Observe that
(7.5)
$$\mathrm{\Sigma }_+^{}=\frac{3}{2}N_1^2(2\mathrm{\Sigma }_+),\mathrm{\Sigma }_{}^{}=\frac{3}{2}N_1^2\mathrm{\Sigma }_{}$$
and
(7.6)
$$\omega ^{}=\frac{3}{2}N_1^2\omega .$$
Consequently, $`\mathrm{\Sigma }_+`$, $`\mathrm{\Sigma }_{}`$ and $`\omega `$ are all monotone so that they converge, both as $`\tau \mathrm{}`$ and as $`\tau \mathrm{}`$. It also follows from (7.5) and (7.6) that the quotients $`(2\mathrm{\Sigma }_+)/\omega `$ and $`\mathrm{\Sigma }_{}/\omega `$ are constant. Thus the orbit in $`\omega \mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-space describes a straight line connecting two points satisfying (7.2). As $`\tau \mathrm{}`$, the solution cannot converge to a point in $`_1_1`$ for the following reason. Assume it does. Since $`\mathrm{\Sigma }_+`$ decreases as $`\tau `$ decreases, see (7.5), we must have $`\mathrm{\Sigma }_+1/2`$ for the entire solution, since $`\mathrm{\Sigma }_+`$ by assumption converges to a value $`1/2`$. But then $`N_1^{}<0`$ for the entire solution by (2.1) and (2.3). Thus $`N_1`$ increases as we go backward, contradicting the fact that $`N_10`$. $`\mathrm{}`$
The next thing we wish to prove is that if a solution has an $`\alpha `$-limit point $`x`$ in the set $`_1`$, and $`N_10`$ for the solution, then we can apply the โKasnerโ map to that point. What we mean by that is that an entire type II orbit with $`x`$ as an $`\omega `$-limit point belongs to the $`\alpha `$-limit set of the original solution. From this one can draw quite strong conclusions. Observe for instance that by (7.1), $`\omega `$ is monotone for a Bianchi VIII solution to (2.1)-(2.3). Thus $`\omega `$ converges as $`\tau \mathrm{}`$ since it is bounded. If the Bianchi VIII solution has an $`\alpha `$-limit point of type I outside the triangle, we can apply the Kasner map to it to obtain $`\alpha `$-limit points with different $`\omega `$. But that is impossible.
###### Lemma 7.2.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ such that $`N_10`$. Then if the solution has an $`\alpha `$-limit point $`x_1`$, the orbit of a type II solution with $`x`$ as an $`\omega `$-limit point belongs to the $`\alpha `$-limit set of the solution.
Proof. The proof is analogous to the proof of Proposition 6.1. $`\mathrm{}`$
Consider a solution such that $`\omega >0`$. We want to exclude the possibility that $`\omega 0`$ as $`\tau \mathrm{}`$. Considering (7.1), we see that the only possibility for $`\omega `$ to decrease is if $`q>2`$. In that context, the following lemma is relevant.
###### Lemma 7.3.
Consider a Bianchi IX solution to (2.1)-(2.3) with $`\gamma =2`$. There is an $`\alpha _0`$ such that if $`\alpha \alpha _0`$ and
$$(N_1N_2N_3)(\tau )\alpha ,$$
then
$$q(\tau )24\alpha ^{1/3}.$$
Proof. By a permutation of the variables, we can assume $`N_1N_2N_3`$ in $`\tau `$. Observe that
$$q23N_1(N_2+N_3)$$
by the constraint (2.3). If $`N_3\alpha ^{1/2}`$ in $`\tau `$, we get $`q26\alpha 4\alpha ^{1/3}`$ if $`\alpha _0`$ is small enough. If $`N_3\alpha ^{1/2}`$ in $`\tau `$, we get
$$N_1N_2\alpha ^{1/2}.$$
Assume, in order to reach a contradiction, $`(N_1N_3)(\tau )\alpha ^{1/3}`$. Then $`N_2(\tau )\alpha ^{2/3}`$, so that $`N_1(\tau )\alpha ^{2/3}`$ and $`N_3(\tau )\alpha ^{1/3}`$. By Lemma 3.3 we get a contradiction if $`\alpha _0`$ is small enough. Thus
$$q23(N_1N_2+N_1N_3)(\tau )3(\alpha ^{1/3}+\alpha ^{1/2})4\alpha ^{1/3}$$
if $`\alpha _0`$ is small enough. $`\mathrm{}`$
For all solutions except those of Bianchi IX type, $`\omega `$ is monotone increasing as $`\tau `$ decreases. Thus, $`\omega `$ is greater than zero on the $`\alpha `$-limit set of any non-vacuum solution which is not of type IX. It turns out that the same is true for a Bianchi IX solution.
###### Lemma 7.4.
Consider a Bianchi IX solution to (2.1)-(2.3) with $`\gamma =2`$ such that $`\omega >0`$. Then there is an $`ฯต>0`$ such that $`\omega (\tau )ฯต`$ for all $`\tau 0`$.
Proof. Assume all the $`N_i`$ are positive. The function
$$\varphi =\frac{(N_1N_2N_3)^{1/3}}{\omega }$$
satisfies $`\varphi ^{}=2\varphi `$. Thus, for $`\tau 0`$,
$$(N_1N_2N_3)^{1/3}(\tau )=\omega (\tau )\varphi (0)e^{2\tau }Ce^{2\tau },$$
because of Lemma 3.3. For $`\tau T0`$, we can thus apply Lemma 7.3, so that for $`\tau T`$,
$$_\tau ^0(q(s)2)๐s=_\tau ^T(q(s)2)๐s+_T^0(q(s)2)๐s4C_\tau ^Te^{2s}๐s+$$
$$+_T^0(q(s)2)๐s2Ce^{2T}+_T^0(q(s)2)๐sC^{}<\mathrm{}.$$
Consequently,
$$\omega (\tau )=\omega (0)\mathrm{exp}(_\tau ^0(q(s)2)๐s)\omega (0)e^C^{},$$
and the lemma follows. $`\mathrm{}`$
The next lemma will be used to prove that $`\omega `$ converges for a Bianchi IX solution.
###### Lemma 7.5.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\omega >0`$. Then there is an $`\alpha >0`$ and a $`T`$ such that
$$|N_1N_2|+|N_2N_3|+|N_3N_1|e^{\alpha \tau }$$
for all $`\tau T`$.
Proof. Consider $`g=|N_2N_3|/\omega `$. Then
$$g^{}=(2\omega ^2+2(1+\mathrm{\Sigma }_+)^2+2\mathrm{\Sigma }_{}^2)g.$$
Since $`\omega (\tau )ฯต`$ for all $`\tau 0`$, we conclude that
$$g(\tau )g(0)\mathrm{exp}(2ฯต^2\tau )$$
so that
$$|(N_2N_3)(\tau )|g(0)\omega (\tau )\mathrm{exp}(2ฯต^2\tau ).$$
There are similar estimates for the other products. By Lemma 3.3, we know that $`\omega `$ is bounded in $`(\mathrm{},0]`$ so that by choosing $`\alpha =ฯต^2`$ and $`T`$ negative enough the lemma follows. $`\mathrm{}`$
###### Corollary 7.1.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\omega >0`$. Then $`(\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},N_1,N_2,N_3)(\mathrm{},0]`$ is contained in a compact set and all the $`\alpha `$-limit points are of type I or II.
###### Lemma 7.6.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\omega >0`$. Then
$$\underset{\tau \mathrm{}}{lim}\omega (\tau )=\omega _0>0.$$
Proof. Since this follows from the monotonicity of $`\omega `$ in all cases except Bianchi IX, see (7.1), we assume that the solution is of type IX. Let $`\tau _k\mathrm{}`$ be a sequence such that $`\omega (\tau _k)\omega _1>0`$. This is possible since $`\omega `$ is constrained to belong to a compact set for $`\tau 0`$ by Lemma 3.3, and since $`\omega `$ is bounded away from zero to the past by Lemma 7.4. Assume $`\omega `$ does not converge to $`\omega _1`$. Then there is a sequence $`s_k\mathrm{}`$ such that $`\omega (s_k)\omega _2`$ where we can assume $`\omega _2>\omega _1`$. We can also assume $`\tau _ks_k`$. Then
$$\omega (s_k)=\mathrm{exp}(_{\tau _k}^{s_k}(q2)๐s)\omega (\tau _k).$$
Since
$$q23(N_1N_2+N_2N_3+N_3N_1)3e^{\alpha \tau }$$
for $`\tau T`$ by Lemma 7.5 and the constraint (2.3), we have, assuming $`s_kT`$,
$$_{\tau _k}^{s_k}(q2)๐s3_{\tau _k}^{s_k}e^{\alpha \tau }๐\tau \frac{3}{\alpha }e^{\alpha s_k}.$$
Thus
$$\omega (s_k)\mathrm{exp}(\frac{3}{\alpha }e^{\alpha s_k})\omega (\tau _k)\omega _1,$$
so that $`\omega _2\omega _1`$ contradicting our assumption. $`\mathrm{}`$
###### Corollary 7.2.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\omega >0`$. Then
$$\underset{\tau \mathrm{}}{lim}N_i(\tau )=0$$
for $`i=1,2,3`$.
Proof. Assume $`N_1`$ does not converge to zero. Then there is a type II $`\alpha `$-limit point with $`N_1`$ and $`\omega `$ non-zero by Corollary 7.1 and Lemma 7.6. If we apply the flow, we get $`\alpha `$-limit points with different $`\omega `$ in contradiction to Lemma 7.6. $`\mathrm{}`$
###### Lemma 7.7.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\omega >0`$. If it has an $`\alpha `$-limit point of type I inside the triangle, the solution converges to that point.
Proof. Let $`x`$ be the limit point. Let $`B`$ be a ball of radius $`ฯต`$ in $`\omega \mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-space, with center given by the $`\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{}`$-coordinates of $`x`$. Let $`\tau _k\mathrm{}`$ be a sequence that yields $`x`$. Assume the solution leaves $`B`$ to the past of every $`\tau _k`$. Then there is a sequence $`s_k\mathrm{}`$, such that the $`\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{}`$-coordinates of the solution evaluated in $`s_k`$ converges to a point on the boundary of $`B`$, $`s_k\tau _k`$, and the $`\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{}`$-coordinates of the solution are contained in $`B`$ during $`[s_k,\tau _k]`$, $`k`$ large enough.
Since all expressions in the $`N_i`$ decay exponentially as $`e^{\alpha \tau }`$, for some $`\alpha >0`$, as long as the $`\omega ,\mathrm{\Sigma }_+,\mathrm{\Sigma }_{}`$-coordinates are in $`B`$ ($`ฯต`$ small enough), we have
$$|\mathrm{\Sigma }_+^{}|+|\mathrm{\Sigma }_{}^{}|+|\omega ^{}|\alpha _ke^{\alpha (\tau \tau _k)}$$
for $`\tau [s_k,\tau _k]`$ where $`\alpha _k0`$. We get
$$|\mathrm{\Sigma }_+(\tau _k)\mathrm{\Sigma }_+(s_k)|\frac{\alpha _k}{\alpha }0,$$
and similarly for $`\mathrm{\Sigma }_{}`$ and $`\omega `$. The assumption that we always leave $`B`$ consequently yields a contradiction. We must thus converge to the given $`\alpha `$-limit point. $`\mathrm{}`$
###### Proposition 7.1.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\omega >0`$. If $`N_i`$ is non-zero for the solution, it converges to a type I point in the complement of $`_i`$ with $`\omega >0`$.
Proof. If there is an $`\alpha `$-limit point on $`_i`$, we can use Lemma 7.2 to obtain a contradiction to Lemma 7.6. If there is an $`\alpha `$-limit point in $`_k`$ and $`N_k`$ is zero for the solution, the solution converges to that point by an argument similar to the one given in the previous lemma. What remains is the possibility that all the $`\alpha `$-limit points are on the $`_k`$. Since $`\omega `$ converges, the possible points projected to the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane are the intersection between a triangle and a circle. Since the $`\alpha `$-limit set is connected, we conclude that the solution must converge to a point on one of the $`_k`$. $`\mathrm{}`$
###### Proposition 7.2.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\omega >0`$. If $`N_i`$ is non-zero for the solution, the solution cannot converge to a point in $`_i`$.
Proof. Assume $`i=1`$. Then $`_i`$ is the subset of (7.2) consisting of points with $`\mathrm{\Sigma }_+=1/2`$ and $`\omega >0`$. Since $`N_2,N_3,N_2N_3,N_2N_1`$ and $`N_3N_1`$ converge to zero faster than $`N_1^2`$, $`\mathrm{\Sigma }_+^{}`$ will in the end be positive, cf. (7.5), so that there is a $`T`$ such that $`\mathrm{\Sigma }_+(\tau )1/2`$ for $`\tau T`$. Since $`N_1`$ will dominate in the end, we can also assume $`q(\tau )<2`$ for $`\tau T`$. By (2.1) we conclude that $`|N_1|`$ increases backward as $`\tau T`$ contradicting Corollary 7.2. $`\mathrm{}`$
Adding up the last two propositions, we conclude that the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-variables of Bianchi VIII and IX solutions converge to a point interior to the triangle of Figure 2, and $`\mathrm{\Omega }`$ to the value then determined by the constraint (2.3). In the Bianchi VI$`\mathrm{I}_0`$ case, a side of the triangle disappears, increasing the set of points to which $`\mathrm{\Sigma }_+,\mathrm{\Sigma }_{}`$ may converge. We sum up the conclusions in Section 19.
## 8. Type I solutions
Consider type I solutions ($`N_i=0`$). The point $`F`$ and the points on the Kasner circle are fixed points. Consider a solution with $`0<\mathrm{\Omega }(\tau _0)<1`$. Using the constraint, we may express the time derivative of $`\mathrm{\Omega }`$ in terms of $`\mathrm{\Omega }`$. Solving the resulting equation yields
$$\underset{\tau \mathrm{}}{lim}\mathrm{\Omega }(\tau )=0,\underset{\tau \mathrm{}}{lim}\mathrm{\Omega }(\tau )=1.$$
By (2.1) $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ moves radially.
###### Proposition 8.1.
For a type I solution, with $`2/3<\gamma <2`$, which is not F, we have
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},\mathrm{\Omega })(\tau )=(\sigma _+/|\sigma |,\sigma _{}/|\sigma |,0),$$
where $`(\sigma _+,\sigma _{})`$ is the initial value of $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$, and $`|\sigma |`$ is the Euclidean norm of the initial value.
## 9. Type II solutions
###### Proposition 9.1.
Consider a type II solution with $`N_1>0`$ and $`2/3<\gamma <2`$. If the initial value for $`\mathrm{\Sigma }_{}`$ is non-zero, the $`\alpha `$-limit set is a point in $`๐ฆ_2๐ฆ_3`$. If the initial value for $`\mathrm{\Sigma }_{}`$ is zero, either the solution is the special point $`P_1^+(II)`$, it is contained in $`_{\mathrm{II}}`$, or
(9.1)
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Omega },\mathrm{\Sigma }_+,N_1)(\tau )=(0,1,0).$$
Proof. Let the initial data be given by $`(\sigma _+,\sigma _{},\mathrm{\Omega }_0)`$. The vacuum case was handled in Proposition 5.1, so we will assume $`\mathrm{\Omega }_0>0`$.
Consider first the case $`\sigma _{}0`$. Compute
$$q2=\frac{3}{2}(2\gamma )\mathrm{\Omega }\frac{3}{2}N_1^2.$$
Thus, $`\mathrm{\Sigma }_{}`$ decreases if it is negative, and increases if it is positive, as we go backward in time, by (2.1). Thus, both $`N_1`$ and $`\mathrm{\Omega }`$ must converge to $`0`$ as $`\tau \mathrm{}`$, since the variables are constrained to belong to a compact set, and because of the monotonicity principle. Since $`\mathrm{\Sigma }_{}`$ is monotonous and the $`\alpha `$-limit set is connected, see Lemma 3.1, $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ must converge to a point, say $`(s_+,s_{})`$ on the Kasner circle. We must have $`s_{}0`$, and
$$2s_+^2+2s_{}^24s_+0,$$
since $`N_1`$ converges to $`0`$. There are two special points in this set, but we may not converge to them, since that would imply $`N_1=0`$ for the entire solution by Proposition 3.1. The first part of the proposition follows.
Consider the case $`\sigma _{}=0`$. There is a fixed point $`P_1^+(II)`$. Eliminating $`\mathrm{\Omega }`$ from (2.1)-(2.3), we are left with the two variables $`N_1`$ and $`\mathrm{\Sigma }_+`$. The linearization has negative eigenvalues at $`P_1^+(II)`$, so that no solution which does not equal $`P_1^+(II)`$ can have it as an $`\alpha `$-limit point, cf. pp. 228-234. There is also a set of solutions converging to the fixed point $`F`$. Consider now the complement of the above. The function
$$Z_7=\frac{N_1^{2m}\mathrm{\Omega }^{1m}}{(1v\mathrm{\Sigma }_+)^2},$$
where $`v=(3\gamma 2)/8`$ and $`m=3v(2\gamma )/8(1v^2)`$, found by Uggla satisfies
$$Z_7^{}=\frac{3(2\gamma )}{1v\mathrm{\Sigma }_+}\frac{1}{1v^2}(\mathrm{\Sigma }_+v)^2Z_7.$$
Apply the monotonicity principle. Let $`G=Z_7`$ and $`U`$ be defined as the subset of $`\mathrm{\Omega }\mathrm{\Sigma }_+N_1`$-space consisting of points different from $`P_1^+(II)`$, which have $`\mathrm{\Omega }>0`$, $`N_1>0`$ and $`|\mathrm{\Sigma }_+|<1`$. Let $`M`$ be defined by the constraint. If $`\mathrm{\Sigma }_+=v`$ then $`Z_7^{}=0`$, but if we are not at $`P_1^+(II)`$, $`\mathrm{\Sigma }_+=v`$ implies $`\mathrm{\Sigma }_+^{}0`$. Thus, $`Gx`$ is strictly monotone as long as $`x`$ is contained in $`UM`$. Since the solution cannot have $`P_1^+(II)`$ as an $`\alpha `$-limit point, we must thus have $`N_1=0`$ or $`\mathrm{\Omega }=0`$ in the $`\alpha `$-limit set. Observe that
(9.2)
$$\mathrm{\Sigma }_+^{}=\frac{3}{2}N_1^2(2\mathrm{\Sigma }_+)\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_+.$$
Thus, if the solution attains a point $`\mathrm{\Sigma }_+0`$, then (9.1) holds. We will now prove that this is the only possibility.
a. Assume we have an $`\alpha `$-limit point with $`N_1>0`$ and $`\mathrm{\Omega }=0`$. Then we may apply the flow to that limit point to get $`\mathrm{\Sigma }_+=1`$ as a limit point, but then the solution must attain $`\mathrm{\Sigma }_+0`$.
b. If $`\mathrm{\Omega }>0`$ but $`N_1=0`$, then we may assume $`\mathrm{\Sigma }_+0`$ since we are not on $`_{\mathrm{II}}`$, cf. Lemma 4.2. Apply the flow to arrive at $`\mathrm{\Sigma }_+=1`$ or $`\mathrm{\Sigma }_+=1`$. The former alternative has been dealt with, and the latter case allows us to construct an $`\alpha `$-limit point with $`N_1>0`$ and $`\mathrm{\Omega }=0`$, since $`N_1`$ increases exponentially, and $`\mathrm{\Omega }`$ decreases exponentially, in a neighbourhood of the point on the Kasner circle with $`\mathrm{\Sigma }_+=1`$, cf. Proposition 6.1.
c. The situation $`\mathrm{\Omega }=N_1=0`$ can be handled as above. $`\mathrm{}`$
We make one more observation that will be relevant in analyzing the regularity of $`_{\mathrm{II}}`$.
###### Lemma 9.1.
The closure of $`_{\mathrm{II}}`$ does not intersect $`๐`$.
Proof. Assume there is a sequence $`x_k_{\mathrm{II}}`$ such that the distance from $`x_k`$ to $`๐`$ goes to zero. We can assume that all the $`x_k`$ have $`N_1>0`$ by choosing a suitable subsequence and then applying the symmetries. We can also assume that $`x_kx๐`$. Since $`\mathrm{\Sigma }_{}=0`$ for all the $`x_k`$ by Proposition 9.1, the same holds for $`x`$. Observe that no element of $`_{\mathrm{II}}`$ can have $`\mathrm{\Sigma }_+0`$, because of (9.2). If $`N_1`$ corresponding to $`x`$ is zero, we then conclude that $`x`$ is defined by $`\mathrm{\Sigma }_+=1`$ and all the other variables zero. Applying the flow to the past to the points $`x_k`$ will then yield a sequence $`y_k_{\mathrm{II}}`$ such that $`y_k`$ converges to a type II vacuum point with $`N_1>0`$ and $`\mathrm{\Sigma }_{}=0`$, cf. the proof of Proposition 6.1. Thus, we can assume that the limit point $`x๐`$ has $`N_1>0`$. Applying the flow to $`x`$ yields the point $`\mathrm{\Sigma }_+=1`$ on the Kasner circle by Proposition 5.1. By the continuity of the flow, we can apply the flow to $`x_k`$ to obtain elements in $`_{\mathrm{II}}`$ with $`\mathrm{\Sigma }_+<0`$ which is impossible. $`\mathrm{}`$
## 10. Type VI$`\mathrm{I}_0`$ solutions
When speaking of Bianchi VI$`\mathrm{I}_0`$ solutions, we will always assume $`N_1=0`$ and $`N_2,N_3>0`$. Consider first the case $`N_2=N_3`$ and $`\mathrm{\Sigma }_{}=0`$
###### Proposition 10.1.
Consider a type VI$`I_0`$ solution with $`N_1=0`$ and $`2/3<\gamma <2`$. If $`N_2=N_3`$ and $`\mathrm{\Sigma }_{}=0`$, one of the following possibilities occurs
1. The solution converges to $`\mathrm{\Sigma }_+=1`$ on the Kasner circle.
2. The solution converges to $`F`$.
3. $`lim_\tau \mathrm{}\mathrm{\Sigma }_+=1,lim_\tau \mathrm{}N_2=n_2>0,lim_\tau \mathrm{}\mathrm{\Omega }=0`$.
Proof. Since
$$\mathrm{\Sigma }_+^{}=\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_+$$
if $`N_2=N_3`$, the conclusions of the lemma follow, except for the statement that $`N_2`$ converges to a non-zero value if $`\mathrm{\Sigma }_+`$ converges to $`1`$. However, $`\mathrm{\Omega }`$ will decay to zero exponentially close to the Kasner circle, and by the constraint, $`1+\mathrm{\Sigma }_+`$ will behave as $`\mathrm{\Omega }`$ close to $`\mathrm{\Sigma }_+=1`$. Thus, $`q+2\mathrm{\Sigma }_+`$ will be integrable. $`\mathrm{}`$
Before we state a proposition concerning the behaviour of generic Bianchi VI$`\mathrm{I}_0`$ solutions, let us give an intuitive picture. Figure 4 shows a simulation with $`\gamma =1`$, where the plus sign represents the starting point, and the star the end point, going backward. $`\mathrm{\Omega }`$ will decay to zero quite rapidly, and the same holds for the product $`N_2N_3`$. In that sense, the solution will asymptotically behave like a sequence of type II vacuum orbits. If both $`N_2`$ and $`N_3`$ are small, and we are close to the section $`๐ฆ_2`$ on the Kasner circle, then $`N_2`$ will increase exponentially, and $`N_3`$ will decay exponentially, yielding in the end roughly a type II orbit with $`N_2>0`$. If this orbit ends in at a point in $`๐ฆ_3`$, then the game begins anew, and we get roughly a type II orbit with $`N_3>0`$. Observe however that if we get close to $`๐ฆ_1`$, there is nothing to make us bounce away, since $`N_1`$ is zero. The simulation illustrates this behaviour. Consider the figure of the solution projected to the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-plane. The three points that appear to be on the Kasner circle are close to $`๐ฆ_2`$, $`๐ฆ_3`$ and $`๐ฆ_1`$ respectively. Observe how this correlates with the graphs of $`N_2`$, $`N_3`$ and $`q`$.
###### Proposition 10.2.
Generic Bianchi VI$`I_0`$ solutions with $`N_1=0`$ and $`2/3<\gamma <2`$ converge to a point in $`๐ฆ_1`$.
We divide the proof into lemmas. First we prove that the past dynamics are contained in a compact set.
###### Lemma 10.1.
For a generic Bianchi VI$`I_0`$ solution with $`N_1=0`$ and $`2/3<\gamma <2`$, $`(N_2,N_3)(\mathrm{},0]`$ is contained in a compact set.
Proof. For a generic solution,
$$Z_1=\frac{\frac{4}{3}\mathrm{\Sigma }_{}^2+(N_2N_3)^2}{N_2N_3}$$
is never zero. Compute
(10.1)
$$Z_1^{}=\frac{16}{3}\frac{\mathrm{\Sigma }_{}^2(1+\mathrm{\Sigma }_+)}{\frac{4}{3}\mathrm{\Sigma }_{}^2+(N_2N_3)^2}Z_1.$$
The proof that the past dynamics are contained in a compact set is as in Rendall . Let $`\tau 0`$. Then
$$Z_1(\tau )Z_1(0),$$
so that
$$(N_2N_3)(\tau )\frac{4}{3Z_1(0)}.$$
Combining this fact with the constraint, we see that all the variables are contained in a compact set during $`(\mathrm{},0]`$. $`\mathrm{}`$
We now prove that $`N_2N_30`$. The reason being the desire to reduce the problem by proving that all the limit points are of type I or II, and then use our knowledge about what happens when we apply the flow to such points.
###### Lemma 10.2.
Generic Bianchi VI$`I_0`$ solutions with $`N_1=0`$ and $`2/3<\gamma <2`$ satisfy
$$\underset{\tau \mathrm{}}{lim}(N_2N_3)(\tau )=0.$$
Proof. Assume the contrary. Then we can use Lemma 10.1 to construct an $`\alpha `$-limit point $`(\omega ,\sigma _+,\sigma _{},0,n_2,n_3)`$ where $`n_2n_3>0`$. We apply the monotonicity principle in order to arrive at a contradiction. With notation as in Lemma 5.1, let $`U`$ be defined by $`N_2>0,N_3>0`$ and $`\mathrm{\Sigma }_{}^2+(N_2N_3)^2>0`$. Let $`G`$ be defined by $`Z_1`$, and $`M`$ by the constraint (2.3). We have to show that $`G`$ evaluated on a solution is strictly monotone as long as the solution is contained in $`UM`$. Consider (10.1). By the constraint (2.3), $`\mathrm{\Sigma }_{}^2+(N_2N_3)^2>0`$ implies $`\mathrm{\Sigma }_+>1`$. Furthermore, $`Z_1>0`$ on $`U`$. If $`Z_1^{}=0`$ in $`UM`$, we thus have $`\mathrm{\Sigma }_{}=0`$, but then $`\mathrm{\Sigma }_{}^{}0`$ since $`\mathrm{\Sigma }_{}^2+(N_2N_3)^2>0`$ and $`N_2+N_3>0`$. The $`\alpha `$-limit point we have constructed cannot belong to $`U`$. On the other hand, $`n_2,n_3>0`$ and since $`Z_1`$ increases as we go backward, $`\sigma _{}^2+(n_2n_3)^2`$ cannot be zero. We have a contradiction. $`\mathrm{}`$
Proof of Proposition 10.2. Compute
(10.2)
$$\mathrm{\Sigma }_+^{}=(22\mathrm{\Omega }2\mathrm{\Sigma }_+^22\mathrm{\Sigma }_{}^2)(1+\mathrm{\Sigma }_+)\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_+$$
by (2.4). Assume we are not on $`๐ซ_{\mathrm{VII}_0}`$ or $`_{\mathrm{VII}_0}`$. Let us first prove that there is an $`\alpha `$-limit point on the Kasner circle. Assume $`F`$ is an $`\alpha `$-limit point. Then we may construct a type I limit point which is not $`F`$, and thus a limit point on the Kasner circle, cf. Lemma 4.2 and Proposition 8.1. By Lemma 10.2, we may then assume that there is a limit point of type I or II, which is not $`P_2^+(II)`$ or $`P_3^+(II)`$, and does not lie in $`_\mathrm{I}`$ or $`_{\mathrm{II}}`$, cf. Lemma 4.1. Thus, we get a limit point on the Kasner circle by Proposition 8.1 and Proposition 9.1.
Next, we prove that there has to be an $`\alpha `$-limit point which lies in the closure of $`๐ฆ_1`$. If the $`\alpha `$-limit point we have constructed is in $`๐ฆ_2`$ or $`๐ฆ_3`$, we can apply the Kasner map according to the remark following Proposition 6.1. After a finite number of Kasner iterates we will end up in the desired set. If the $`\alpha `$-limit point we obtained has $`\mathrm{\Sigma }_+=1`$, we may construct a limit point with $`1+\mathrm{\Sigma }_+=ฯต>0`$ by Proposition 3.1. We can also assume that $`\mathrm{\Omega }=0`$ for this point, since $`\mathrm{\Omega }`$ decays exponentially going backward when $`\mathrm{\Sigma }_+`$ is close to $`1`$. By Lemma 10.2, this limit point will be a type I or II vacuum point, and by applying the flow we get a non special limit point on the Kasner circle. As above, we then get an $`\alpha `$-limit point in the desired set. Let the $`\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$-variables of one $`\alpha `$-limit point in the closure of $`๐ฆ_1`$ be $`(\sigma _+,\sigma _{})`$.
By (10.2), we conclude that once $`\mathrm{\Sigma }_+`$ has become greater than $`0`$, it becomes monotone so that it has to converge. Moreover, we see by the same equation that $`\mathrm{\Omega }`$ then has to converge to zero, and $`\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2`$ has to converge to $`1`$. Since the $`\alpha `$-limit set is connected, by Lemma 3.1 and Lemma 10.1, we conclude that $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ has to converge to $`(\sigma _+,\sigma _{})`$. By Proposition 3.1, $`(\sigma _+,\sigma _{})`$ cannot equal $`(1/2,\pm \sqrt{3}/2)`$, since otherwise $`N_2`$ or $`N_3`$ would be zero for the entire solution. Consequently, $`\sigma _+>1/2`$, and we conclude that $`N_2`$ and $`N_3`$ have to converge to zero. The proposition follows. $`\mathrm{}`$
## 11. Taub type IX solutions
Consider the Taub type solutions: $`\mathrm{\Sigma }_{}=0`$ and $`N_2=N_3`$. We prove that except for the cases when the solution belongs to $`_{\mathrm{IX}}`$ or $`๐ซ_{\mathrm{IX}}`$, $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ converges to $`(1,0)`$.
###### Lemma 11.1.
Consider a type IX solution with $`\mathrm{\Sigma }_{}=0`$, $`N_2=N_3`$ and $`2/3<\gamma <2`$. Then $`\mathrm{\Sigma }_+(\tau _0)0`$ and $`\mathrm{\Omega }(\tau _0)<1`$ imply
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},N_1,N_2,N_3)(\tau )=(0,1,0,0,n_2,n_2),$$
where $`0<n_2<\mathrm{}`$.
Proof. We prove that the flow will take us to the boundary of the parabola $`\mathrm{\Omega }+\mathrm{\Sigma }_+^2=1`$ with $`\mathrm{\Sigma }_+<0`$, and that we will then slide down the side on the outside to reach $`\mathrm{\Sigma }_+=1`$, see Figure 5. The plus sign in the figure represents the starting point, and the star the end point.
1. Let us first assume $`\mathrm{\Sigma }_+(\tau _0)0`$, $`\mathrm{\Omega }(\tau _0)<1`$ and $`\mathrm{\Omega }(\tau _0)+\mathrm{\Sigma }_+^2(\tau _0)1`$. Consider
$$๐=\{\tau \tau _0:t[\tau ,\tau _0]\mathrm{\Sigma }_+(t)0,\mathrm{\Omega }(t)\mathrm{\Omega }(\tau _0),\mathrm{\Omega }(t)+\mathrm{\Sigma }_+^2(t)1\}.$$
We prove that $`๐`$ is not bounded from below. Assume the contrary. Let $`t`$ be the infimum of $`๐`$, which exists since $`๐`$ is non-empty and bounded from below. Since $`t๐`$, $`\mathrm{\Sigma }_+(t)<0`$. Let $`t^{}<t`$ be such that $`\mathrm{\Sigma }_+<0`$ in $`[t^{},t]`$. Observe that
(11.1)
$$\mathrm{\Omega }^{}=[(3\gamma 2)(\mathrm{\Omega }+\mathrm{\Sigma }_+^21)+3(2\gamma )\mathrm{\Sigma }_+^2]\mathrm{\Omega }.$$
By the constraint,
(11.2)
$$\mathrm{\Omega }+\mathrm{\Sigma }_+^21=\frac{3}{4}N_1^2(4\frac{N_2}{N_1}1).$$
Since $`\mathrm{\Sigma }_+<0`$ in $`[t^{},t]`$, $`N_2/N_1`$ increases as we go backward in that interval, because of
$$(\frac{N_2}{N_1})^{}=6\mathrm{\Sigma }_+\frac{N_2}{N_1}.$$
Consequently $`\mathrm{\Omega }+\mathrm{\Sigma }_+^21`$ in $`[t^{},t]`$, by (11.2), so that $`\mathrm{\Omega }`$ decreases in the interval by (11.1). Thus $`t^{}๐`$, contradicting the fact that $`t`$ is the infimum of $`๐`$.
Let $`\tau \tau _0`$. Then $`\mathrm{\Sigma }_+(\tau )\sqrt{1\mathrm{\Omega }(\tau _0)}`$. By (11.1), we then conclude $`\mathrm{\Omega }0`$. By (2.1), we also conclude that $`N_1N_20`$ and $`N_10`$. By (11.2), we have $`\mathrm{\Sigma }_+1`$. Using the constraint (11.2) and (2.2), we conclude that $`q+2\mathrm{\Sigma }_+`$ is integrable, so that $`N_2=N_3`$ will converge to a finite non-zero value.
2. Assume now $`\mathrm{\Sigma }_+(\tau _0)0`$, $`\mathrm{\Omega }(\tau _0)<1`$ and $`\mathrm{\Omega }(\tau _0)+\mathrm{\Sigma }_+^2(\tau _0)<1`$. Observe that
(11.3)
$$\mathrm{\Sigma }_+^{}=(1\mathrm{\Omega }\mathrm{\Sigma }_+^2)(42\mathrm{\Sigma }_+)\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_++9N_1N_2.$$
As long as $`\mathrm{\Omega }+\mathrm{\Sigma }_+^2<1`$, $`\mathrm{\Sigma }_+`$ decreases as we go backward in time by (11.3). Then $`N_2/N_1`$ will increase exponentially until $`\mathrm{\Omega }+\mathrm{\Sigma }_+^2=1`$, by the constraint, and $`\mathrm{\Sigma }_+<0`$. $`\mathrm{}`$
###### Lemma 11.2.
Consider a type IX solution with $`\mathrm{\Sigma }_{}=0`$, $`N_2=N_3`$ and $`2/3<\gamma <2`$. It is contained in a compact set for $`\tau 0`$ and $`N_1N_20`$.
Proof. Note that $`N_1`$ must be bounded for $`\tau 0`$, as follows from Lemma 3.3, the fact that $`N_2=N_3`$, and the fact that $`N_1N_2N_3`$ decreases backward in time. To prove the first statement, assume the contrary. Then there is a sequence $`\tau _k\mathrm{}`$ such that $`N_2(\tau _k)\mathrm{}`$. We can assume $`N_2^{}(\tau _k)0`$, and thus
(11.4)
$$\frac{1}{2}(3\gamma 2)\mathrm{\Omega }+2\mathrm{\Sigma }_+^2+2\mathrm{\Sigma }_+0$$
in $`\tau _k`$. Since $`N_1N_2^2`$ is decreasing as we go backward, $`N_1`$ and $`N_1N_2`$ evaluated at $`\tau _k`$ must go to zero. Thus $`\mathrm{\Omega }+\mathrm{\Sigma }_+^21`$ will become arbitrarily small in $`\tau _k`$ by (11.2). If $`\mathrm{\Omega }(\tau _k)1`$ for all $`k`$, we get
$$\mathrm{\Sigma }_+(\tau _k)\frac{1}{4}(3\gamma 2)$$
by (11.4), so that
$$\mathrm{\Sigma }_+^2(\tau _k)+\mathrm{\Omega }(\tau _k)1+\frac{1}{16}(3\gamma 2)^2,$$
which is a contradiction. In other words, there is a $`k`$ such that $`\mathrm{\Sigma }_+(\tau _k)0`$, by (11.4), and $`\mathrm{\Omega }(\tau _k)<1`$. We can then use Lemma 11.1 to arrive at a contradiction to the assumption that the solution is not contained in a compact set.
To prove the second part of the lemma, observe that $`N_1N_2^2`$ converges to zero, as follows from the existence of an $`\alpha `$\- limit point and Lemma 5.2. Thus
$$N_1N_2=N_1^{1/2}[N_1N_2^2]^{1/2}C[N_1N_2^2]^{1/2}0.$$
$`\mathrm{}`$
###### Proposition 11.1.
For a type IX solution with $`\mathrm{\Sigma }_{}=0`$, $`N_2=N_3`$ and $`2/3<\gamma <2`$, either the solution is contained in $`_{\mathrm{IX}}`$ or $`๐ซ_{\mathrm{IX}}`$, or
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{},N_1,N_2,N_3)(\tau )=(0,1,0,0,n_2,n_2)$$
where $`0<n_2<\mathrm{}`$.
Remark. Compare with Proposition 3.1. Observe also that when $`\mathrm{\Sigma }_+`$ for the solution converges to $`1`$, we approach $`\mathrm{\Sigma }_+=1,\mathrm{\Omega }=0`$ from outside the parabola $`\mathrm{\Omega }+\mathrm{\Sigma }_+^2=1`$, as follows from the proof of Lemma 11.1.
Proof. Consider a solution which is not contained in $`_{\mathrm{IX}}`$ or $`๐ซ_{\mathrm{IX}}`$. By Lemma 11.2, there is an $`\alpha `$-limit point with $`N_1N_2=0`$. We can assume it is not $`P_1^+(II)`$. We have the following possibilities.
1. It is contained in $`_\mathrm{I}_{\mathrm{II}}_{\mathrm{VII}_0}`$. Then $`F`$ is an $`\alpha `$-limit point. Since the solution is not contained in $`_{\mathrm{IX}}`$, we get a type I limit point which is not $`F`$, by Lemma 4.2, and thus either $`\mathrm{\Sigma }_+=1`$ or $`\mathrm{\Sigma }_+=1`$ as limit points, by Proposition 8.1. The first alternative implies convergence to $`\mathrm{\Sigma }_+=1`$, by Lemma 11.1. If we have a type I $`\alpha `$-limit point with $`\mathrm{\Sigma }_+=1`$, we can apply the Kasner map by Proposition 6.1 in order to obtain a type I limit point with $`\mathrm{\Sigma }_+=1`$.
2. The limit point is of type I. This possibility can be dealt with as above.
3. It is of type II. We can assume that it is not $`P_1^+(II)`$, by Lemma 4.1, and that it is not contained in $`_{\mathrm{II}}`$. Thus we get $`\mathrm{\Sigma }_+=1`$ on the Kasner circle as an $`\alpha `$-limit point, by Proposition 9.1, and thus as above convergence to $`\mathrm{\Sigma }_+=1`$.
4. The limit point is of type VI$`I_0`$. We can assume $`\mathrm{\Sigma }_+0`$. If $`\mathrm{\Sigma }_+<0`$, we can apply Lemma 11.1 again, and if $`\mathrm{\Sigma }_+>0`$, we get $`\mathrm{\Sigma }_+=1`$ on the Kasner circle as an $`\alpha `$-limit point, by Proposition 10.1, a case which can be dealt with as above. $`\mathrm{}`$
## 12. Oscillatory behaviour
It will be necessary to consider Bianchi IX solutions to (2.1)-(2.3) under circumstances such that the behaviour is oscillatory. This section provides the technical tools needed.
Let $`g`$ be a function,
(12.1)
$$A=\left(\begin{array}{cc}\hfill 0& \hfill g\\ \hfill g& \hfill 0\end{array}\right),$$
and $`\stackrel{~}{๐ฑ}=(\stackrel{~}{x},\stackrel{~}{y})^t`$ satisfy
$$\stackrel{~}{๐ฑ}^{}=A\stackrel{~}{๐ฑ}+ฯต,$$
where $`ฯต`$ is some vector valued function.
###### Lemma 12.1.
Let $`\varphi _0`$ be such that $`(\mathrm{sin}(\varphi _0),\mathrm{cos}(\varphi _0))`$ and $`(\stackrel{~}{x}(\tau _0),\stackrel{~}{y}(\tau _0))`$ are parallel. Define
(12.2)
$$\xi (\tau )=_{\tau _0}^\tau g(s)๐s+\varphi _0$$
and
(12.3)
$$๐ฑ(\tau )=\left(\begin{array}{c}x(\tau )\\ y(\tau )\end{array}\right)=\left(\begin{array}{c}\mathrm{sin}\xi (\tau )\\ \mathrm{cos}\xi (\tau )\end{array}\right).$$
Then
(12.4)
$$\stackrel{~}{๐ฑ}(\tau )๐ฑ(\tau )|1(\stackrel{~}{x}^2(\tau _0)+\stackrel{~}{y}^2(\tau _0))^{1/2}|+|_{\tau _0}^\tau ฯต(s)ds|.$$
Proof. Let
$$\mathrm{\Phi }=\left(\begin{array}{cc}\hfill y& \hfill x\\ \hfill x& \hfill y\end{array}\right).$$
We have $`[A,\mathrm{\Phi }]=0`$, $`\mathrm{\Phi }^{}=A\mathrm{\Phi }`$ and $`๐ฑ^{}=A๐ฑ`$. We get
$$(\mathrm{\Phi }(\stackrel{~}{๐ฑ}๐ฑ))^{}=A\mathrm{\Phi }(\stackrel{~}{๐ฑ}๐ฑ)+\mathrm{\Phi }(A(\stackrel{~}{๐ฑ}๐ฑ)+ฯต)=\mathrm{\Phi }ฯต.$$
Thus
$$(\stackrel{~}{๐ฑ}๐ฑ)(\tau )=\mathrm{\Phi }^1(\tau )\mathrm{\Phi }(\tau _0)(\stackrel{~}{๐ฑ}๐ฑ)(\tau _0)+\mathrm{\Phi }^1(\tau )_{\tau _0}^\tau \mathrm{\Phi }(s)ฯต(s)๐s.$$
But $`\mathrm{\Phi }`$ takes values in $`SO(2)`$ and the lemma follows. $`\mathrm{}`$
In order to prove the existence of an $`\alpha `$-limit point for Bianchi IX solutions, and that, generically, there is a limit point on the Kasner circle, we need the following lemma.
###### Lemma 12.2.
Consider a Bianchi IX solution with $`2/3<\gamma <2`$. Assume there is a sequence $`\tau _k\mathrm{}`$ such that $`q(\tau _k)0`$, and $`N_2(\tau _k),N_3(\tau _k)\mathrm{}`$, then for each $`T`$, there is a $`\tau T`$ such that $`\mathrm{\Sigma }_+(\tau )0`$.
Proof. Observe that by (2.4), $`q=0`$ and $`N_2+N_3N_1`$ implies $`\mathrm{\Sigma }_+^{}2`$. However, the only term appearing in the constraint which does not go to zero in $`\tau _k`$ is $`(N_2N_3)^2`$, since the product $`N_1N_2N_3`$ decreases as we go backward. Thus $`|\mathrm{\Sigma }_{}^{}(\tau _k)|\mathrm{}`$, and the behaviour is oscillatory. It is clear that $`\mathrm{\Sigma }_+^{}`$ could become positive during the oscillations, but only when $`|\mathrm{\Sigma }_{}|`$ is big, so that we on the whole should move in the positive direction.
Assume there is a $`T`$ such that $`\mathrm{\Sigma }_+(\tau )<0`$ for all $`\tau T`$.
We begin by examining the behaviour of different expressions in the sets
$$๐_k=_{n=k}^{\mathrm{}}[\tau _n1,\tau _n]$$
and
$$๐=_{n=1}^{\mathrm{}}[\tau _n1,\tau _n].$$
Observe that by the fact that $`(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ are constrained to belong to a compact set during $`(\mathrm{},0]`$, according to Lemma 3.3, $`N_2`$ and $`N_3`$ go to infinity uniformly in $`๐`$ (by which we will mean the following):
$$MK:kKN_i(\tau )M\tau ๐_k,i=2,3.$$
Thus $`N_1`$ and $`N_1(N_2+N_3)`$ go to zero uniformly in $`๐`$. By (2.1), $`\mathrm{\Omega }`$ also converges to zero uniformly in $`๐`$. Due to the constraint, we get a bound on $`\mathrm{\Sigma }_{}^2+\frac{3}{4}(N_2N_3)^2`$ in $`๐`$. Consider (2.4). The last two terms go to zero uniformly. If the first term is not negative, $`1\mathrm{\Omega }\mathrm{\Sigma }_+^2\mathrm{\Sigma }_{}^20`$. By the constraint, it will then be bounded by an expression that converges to zero uniformly in $`๐`$. Thus, for every $`\delta >0`$ there is a $`K`$ such that $`kK`$ implies $`\mathrm{\Sigma }_+^{}\delta `$ in $`๐_k`$. Combining this with the fact that $`q(\tau _k)0`$, and the assumption that $`\mathrm{\Sigma }_+(\tau )<0`$ for $`\tau T`$, we conclude that $`\mathrm{\Sigma }_+`$ converges uniformly to zero in $`๐`$.
Next, we use Lemma 12.1 in order to approximate the oscillatory behaviour. Define the functions
$`\stackrel{~}{x}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Sigma }_{}}{(1\mathrm{\Sigma }_+^2)^{1/2}}}`$
$`\stackrel{~}{y}`$ $`=`$ $`{\displaystyle \frac{\sqrt{3}}{2}}{\displaystyle \frac{N_2N_3}{(1\mathrm{\Sigma }_+^2)^{1/2}}}.`$
We can apply Lemma 12.1 with
$$g=3(N_2+N_3)2(1+\mathrm{\Sigma }_+)\stackrel{~}{x}\stackrel{~}{y}=g_1+g_2$$
and $`ฯต_x`$, $`ฯต_y`$ given by (15.5) and (15.6), cf. Lemma 15.1. By the above, we conclude that $`\stackrel{~}{x}`$ and $`\stackrel{~}{y}`$ are uniformly bounded on $`๐_k`$, if $`k`$ is great enough, and that $`ฯต`$ converges to zero uniformly on $`๐`$. Let $`๐ฑ_k`$ be the expression given by Lemma 12.1, with $`\tau _0`$ replaced by $`\tau _k`$ and $`\varphi _0`$ by a suitable $`\varphi _k`$. Let $`\delta >0`$. By the above and $`q(\tau _k)0`$, we get
(12.5)
$$(\stackrel{~}{๐ฑ}๐ฑ_k)(\tau )\delta ,$$
if $`\tau [\tau _k1,\tau _k]`$, and $`k`$ is great enough. In $`[\tau _k1,\tau _k]`$, we thus have
(12.6)
$$\mathrm{\Sigma }_+^{}=2+2x_k^2(1\mathrm{\Sigma }_+^2)+\rho _k,$$
where the error $`\rho _k`$ can be assumed to be arbitrarily small by choosing $`k`$ great enough, cf. (2.4).
Let
$$\xi _k(\tau )=_{\tau _k}^\tau g(s)๐s+\varphi _k$$
be as in (12.2). Since $`N_2+N_3`$ goes to infinity uniformly, $`[\tau _k1,\tau _k]`$ can be assumed to contain an arbitrary number of periods of $`\xi _k`$, if $`k`$ is great enough. Thus, we can assume the existence of $`\tau _{1,k},\tau _{2,k}[\tau _k1,\tau _k]`$, such that $`\tau _{2,k}\tau _{1,k}1/2`$ and $`\xi _k(\tau _{1,k})\xi _k(\tau _{2,k})`$ is an integer multiple of $`\pi `$. Let $`[\tau _1,\tau _2][\tau _{1,k},\tau _{2,k}]`$ satisfy $`\xi _k(\tau _1)\xi _k(\tau _2)=\pi `$. We can assume $`\tau _2\tau _1`$ to be arbitrarily small by choosing $`k`$ great enough. Considering (2.1), and using the fact that $`q`$ is bounded, we conclude that $`N_2+N_3`$ cannot change by more than a factor arbitrarily close to one during $`[\tau _1,\tau _2]`$. Since the expression involving $`N_2+N_3`$ dominates $`g`$, we conclude that
$$\frac{3}{4}g(\tau _{\mathrm{max}})g(\tau _{\mathrm{min}}),$$
where $`\tau _{\mathrm{max}}`$ and $`\tau _{\mathrm{min}}`$ correspond to the maximum and the minimum of $`g`$ in $`[\tau _1,\tau _2]`$. Estimate
$$_{\tau _1}^{\tau _2}2x_k^2(1\mathrm{\Sigma }_+^2)๐s=_{\xi _k(\tau _1)}^{\xi _k(\tau _2)}\frac{2x_k^2(1\mathrm{\Sigma }_+^2)}{g}๐\eta =_{\xi _k(\tau _1)\pi }^{\xi _k(\tau _1)}\frac{2x_k^2(1\mathrm{\Sigma }_+^2)}{g}๐\eta $$
$$\frac{1}{g(\tau _{\mathrm{min}})}_{\xi _k(\tau _1)\pi }^{\xi _k(\tau _1)}2\mathrm{sin}^2(\eta )๐\eta =\frac{\pi }{g(\tau _{\mathrm{min}})}.$$
We get
$$\tau _2\tau _1=_{\xi _k(\tau _1)}^{\xi _k(\tau _2)}\frac{1}{g}๐\eta \frac{\pi }{g(\tau _{\mathrm{max}})}\frac{3}{4}_{\tau _1}^{\tau _2}2x_k^2(1\mathrm{\Sigma }_+^2)๐s.$$
Consequently, (12.6) yields
$$\mathrm{\Sigma }_+(\tau _2)\mathrm{\Sigma }_+(\tau _1)=2(\tau _2\tau _1)+_{\tau _1}^{\tau _2}2x_k^2(1\mathrm{\Sigma }_+^2)๐s+_{\tau _1}^{\tau _2}\rho _k๐\tau $$
$$\frac{2}{3}(\tau _2\tau _1)+_{\tau _1}^{\tau _2}\rho _k๐\tau .$$
Since $`\xi _k(\tau _{1,k})\xi _k(\tau _{2,k})`$ corresponds to an integer multiple of $`\pi `$, we conclude that
$$\mathrm{\Sigma }_+(\tau _{2,k})\mathrm{\Sigma }_+(\tau _{1,k})\frac{2}{3}(\tau _{2,k}\tau _{1,k})+_{\tau _{1,k}}^{\tau _{2,k}}\rho _k๐\tau \frac{1}{3}+_{\tau _{1,k}}^{\tau _{2,k}}\rho _k๐\tau .$$
However, the expressions on the far left can be assumed to be arbitrarily small, and the integral of $`\rho _k`$ can be assumed to be arbitrarily small. We have a contradiction. $`\mathrm{}`$
## 13. Bianchi IX solutions
We first prove that there is an $`\alpha `$-limit point. If we assume that there is no $`\alpha `$-limit point, we get the conclusion that the Euclidean norm $`N`$ of the vector $`(N_1,N_2,N_3)`$ has to converge to infinity, since $`(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ is constrained to belong to a compact set to the past by Lemma 3.3. In fact, Lemma 3.3 yields more; it implies that two $`N_i`$ have to be large at any given time. Since the product $`N_1N_2N_3`$ decays as we go backward, the third $`N_i`$ has to be small. Sooner or later, the two $`N_i`$ which are large and the one which is small have to be fixed, since a โchanging of rolesโ would require two $`N_i`$ to be small, and thereby also the third by Lemma 3.3, contradicting the fact that $`N\mathrm{}`$. Therefore, one can assume that two $`N_i`$ converge to infinity, and that the third converges to zero. More precisely we have.
###### Lemma 13.1.
Consider a Bianchi IX solution. If $`N\mathrm{}`$, we can, by applying the symmetries to the equations, assume that $`N_2,N_3\mathrm{}`$ and $`N_1,N_1(N_2+N_3)0`$.
Proof. As in the vacuum case, see . $`\mathrm{}`$
###### Lemma 13.2.
A Bianchi IX solution with $`2/3<\gamma <2`$ has an $`\alpha `$-limit point.
Proof. If the solution is of Taub type, we already know that it is true so assume not. We assume $`N_2,N_3\mathrm{}`$, since if this does not occur, there is an $`\alpha `$-limit point by Lemma 3.3 and Lemma 13.1. By (2.4) we have $`\mathrm{\Sigma }_+^{}<0`$ if $`\mathrm{\Sigma }_+=0`$ using the constraint (assuming $`N_2+N_3>3N_1`$). Thus, there is a $`T`$ such that if $`\mathrm{\Sigma }_+`$ attains zero in $`\tau T`$, it will be non-negative to the past, and thus $`N_2N_3`$ will be bounded to the past since $`\mathrm{\Sigma }_+`$ has to be negative for the product to grow. If there is a sequence $`\tau _k\mathrm{}`$ such that $`q(\tau _k)0`$, we can apply Lemma 12.2 to arrive at a contradiction. Thus there is an $`S`$ such that
(13.1)
$$q(\tau )ฯต>0$$
for all $`\tau S`$.
Consider
(13.2)
$$Z_1=\frac{\frac{4}{3}\mathrm{\Sigma }_{}^2+(N_2N_3)^2}{N_2N_3}.$$
The reason we consider this function is that the derivative is in a sense almost negative, so that it almost increases as we go backward. On the other hand, it converges to zero as $`\tau \mathrm{}`$ by our assumptions. The lemma follows from the resulting contradiction. We have
(13.3)
$$Z_1^{}=\frac{h}{N_2N_3}=\frac{\frac{16}{3}\mathrm{\Sigma }_{}^2(1+\mathrm{\Sigma }_+)+4\sqrt{3}\mathrm{\Sigma }_{}(N_2N_3)N_1}{N_2N_3}.$$
Letting
$$f=\frac{4}{3}\mathrm{\Sigma }_{}^2+(N_2N_3)^2,$$
we have, using the constraint,
$$h4\mathrm{\Sigma }_{}^2N_1(N_2+N_3)+2\sqrt{3}N_1fN_1N_2N_3f$$
for, say, $`\tau T^{}S`$. Thus
(13.4)
$$Z_1^{}N_1N_2N_3Z_1$$
for all $`\tau T^{}`$. Since $`qฯต>0`$ for all $`\tau T^{}S`$ by (13.1), we get
$$(N_1N_2N_3)(\tau )(N_1N_2N_3)(T^{})\mathrm{exp}[3ฯต(\tau T^{})]$$
for $`\tau T^{}`$. Inserting this inequality in (13.4) we can integrate to obtain
$$Z_1(\tau )Z_1(T^{})\mathrm{exp}(\frac{1}{3ฯต}(N_1N_2N_3)(T^{}))>0$$
for $`\tau T^{}`$. But $`Z_1(\tau )0`$ as $`\tau \mathrm{}`$ by our assumption, and we have a contradiction. $`\mathrm{}`$
###### Corollary 13.1.
Consider a Bianchi IX solution with $`2/3<\gamma <2`$. For all $`ฯต>0`$, there is a $`T`$ such that
$$\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^21+ฯต$$
for all $`\tau T`$. Furthermore
$$\underset{\tau \mathrm{}}{lim}(N_1N_2N_3)(\tau )=0.$$
Proof. As in the vacuum case, see . The second part follows from Lemma 5.2 and Lemma 13.2. $`\mathrm{}`$
###### Proposition 13.1.
A generic Bianchi IX solution with $`2/3<\gamma <2`$ has an $`\alpha `$-limit point on the Kasner circle.
Proof. Observe that by Lemma 13.2 and Corollary 13.1, there is an $`\alpha `$-limit point of type I, II or VI$`\mathrm{I}_0`$.
1. First we prove that we can assume the $`\alpha `$-limit point to be a type VI$`I_0`$ point with $`N_1=0,0<N_2=N_3,\mathrm{\Omega }=0,\mathrm{\Sigma }_{}=0`$ and $`\mathrm{\Sigma }_+=1`$.
a. If there is an $`\alpha `$-limit point in $`_\mathrm{I}`$, $`_{\mathrm{II}}`$ or $`_{\mathrm{VII}_0}`$, $`F`$ is a limit point, but then there is an $`\alpha `$-limit point on the Kasner circle, by Lemma 4.2 and Proposition 8.1.
b. Assume there is an $`\alpha `$-limit point in $`๐ซ_{\mathrm{VII}_0}`$, or that one of $`P_i^+(II)`$ is an $`\alpha `$-limit point. Then there is a limit point of type II which is not $`P_i^+(II)`$, by Lemma 4.1, and we can assume it does not belong to $`_{\mathrm{II}}`$. We thus get an $`\alpha `$-limit point on the Kasner circle by Proposition 9.1.
c. Consider the complement of the above. We have an $`\alpha `$-limit point of type I, II or VI$`\mathrm{I}_0`$ which is generic or possibly of Taub type. If the limit point is of type I or II, we get an $`\alpha `$-limit point on the Kasner circle by Proposition 8.1 and Proposition 9.1. If the limit point is a non-Taub type VI$`\mathrm{I}_0`$ point, we get an $`\alpha `$-limit point on the Kasner circle by Proposition 10.2. Assume it is of Taub type with $`\mathrm{\Sigma }_{}=0`$, $`N_2=N_3`$. By Proposition 10.1, we can assume that we have an $`\alpha `$-limit point of the type mentioned.
2. We construct an $`\alpha `$-limit point on the Kasner circle given an $`\alpha `$-limit point as in 1. Since the solution is not of Taub type, we must leave a neighbourhood of the point $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(1,0)`$. If $`N_2`$ and $`N_3`$ evaluated at the times we leave do not go to infinity, we are done. The reason is that we can choose the neighbourhood to be so small that $`\mathrm{\Omega }`$ and $`N_1`$ decrease exponentially in it, see (2.1). If $`N_2(t_k)`$ or $`N_3(t_k)`$ is bounded, we get a vacuum Bianchi VI$`\mathrm{I}_0`$ $`\alpha `$-limit point which is not of Taub-type by choosing a suitable subsequence (if we get a type I or II point we are done, see the above arguments). By Proposition 10.2, we then get an $`\alpha `$-limit point on the Kasner circle. Thus, we can assume the existence of a sequence $`t_k\mathrm{}`$ such that $`N_2(t_k)`$ and $`N_3(t_k)`$ go to infinity.
There are two problems we have to confront. First of all $`N_2`$ and $`N_3`$ have to decay from their values in $`t_k`$ in order for us to get an $`\alpha `$-limit point. Secondly, and more importantly, we need to see to it that we do not get an $`\alpha `$-limit point of the same type we started with. Let us divide the situation into two cases.
a. Assume that for each $`t_k`$ there is an $`s_kt_k`$ such that $`\mathrm{\Sigma }_+(s_k)=0`$. Observe that when $`\mathrm{\Sigma }_+=0`$, we have
$$\mathrm{\Sigma }_+^{}\frac{1}{2}N_1(9N_13N_23N_3)$$
by the constraint (2.3), and (2.4). Thus, we can assume that we have $`3N_1N_2+N_3`$ in $`s_k`$, since there is an $`\alpha `$-limit point with $`\mathrm{\Sigma }_+=1`$. Thus there must be an $`r_kt_k`$ such that, at $`r_k`$, either $`N_1=N_2<N_3`$, $`N_1=N_3<N_2`$ or $`N_1<N_2`$, $`N_1<N_3`$ and $`3N_1N_2+N_3`$. One of these possibilities must occur an infinite number of times. The first two possibilities yield a type I or II limit point, and the last a type I limit point because, of the fact that $`N_1N_2N_30`$ and Lemma 3.3. As above, we get an $`\alpha `$-limit point on the Kasner circle.
b. Assume there is a $`T`$ such that $`\mathrm{\Sigma }_+(\tau )<0`$ for all $`\tau T`$. Then $`N_10`$, since $`N_1(t_k)0`$, and $`\mathrm{\Sigma }_+<0`$ implies that $`N_1`$ is monotone. Assume there is a sequence $`\tau _k\mathrm{}`$ such that $`N_2`$ or $`N_3`$ evaluated at it goes to zero. Then we get an $`\alpha `$ limit point of type I or II, a situation we may deal with as above. Thus we may assume $`N_iฯต>0`$, $`i=2,3`$ to the past of $`T`$. Similarly to the proof of the existence of an $`\alpha `$-limit point, we have
$$Z_1^{}c_ฯตN_1N_2N_3Z_1.$$
If there is an $`S`$ and a $`\xi >0`$ such that $`q(\tau )\xi >0`$ for all $`\tau S`$, we get a contradiction as in the proof of Lemma 13.2, since $`(N_2N_3)(t_k)\mathrm{}`$. Thus there exists a sequence $`\tau _k\mathrm{}`$ such that $`q(\tau _k)0`$. If $`N_2(\tau _k)`$ or $`N_3(\tau _k)`$ contains a bounded subsequence, we may refer to possibilities already handled. By Lemma 12.2, we get $`\mathrm{\Sigma }_+0`$, a contradiction. $`\mathrm{}`$
## 14. Control over the density parameter
The idea behind the main argument is to use the existence of an $`\alpha `$-limit point on the Kasner circle to obtain a contradiction to the assumption that the solution does not converge to the closure of the set of vacuum type II points. The function
$$d=\mathrm{\Omega }+N_1N_2+N_2N_3+N_3N_1$$
is a measure of the distance from the attractor. We can consider $`d`$ to be a function of $`\tau `$, if we evaluate it at a generic Bianchi IX solution. If $`\tau _k\mathrm{}`$ yields the $`\alpha `$-limit point on the Kasner circle, then $`d(\tau _k)0`$. If $`d`$ does not converge to zero, then it must grow from an arbitrarily small value up to some fixed number, say $`\delta >0`$, as we go backward. In the contradiction argument, it is convenient to know that the growth occurs only in the sum of products of the $`N_i`$, and that during the growth one can assume $`\mathrm{\Omega }`$ to be arbitrarily small. The following proposition achieves this goal, assuming $`\delta `$ is small enough, which is not a restriction. The proof is to be found at the end of this section.
###### Proposition 14.1.
Consider a Bianchi IX solution with $`2/3<\gamma <2`$. There exists an $`ฯต>0`$ such that if
(14.1)
$$N_1N_2+N_2N_3+N_1N_3ฯต$$
in $`[\tau _1,\tau _2]`$, then
$$\mathrm{\Omega }c_\gamma \mathrm{\Omega }(\tau _2)$$
in $`[\tau _1,\tau _2]`$ if $`\mathrm{\Omega }(\tau _2)ฯต`$. Here $`c_\gamma >0`$ only depends on $`\gamma `$.
The idea of the proof is the following. If the sum of product of the $`N_i`$ and $`\mathrm{\Omega }`$ are small, the solution should behave in the following way. If all the $`N_i`$ are small, then we are close to the Kasner circle and $`\mathrm{\Omega }`$ decays exponentially. One of the $`N_i`$ may become large alone, and then $`\mathrm{\Omega }`$ increases, but it can only be large for a short period of time. After that it must decay until some other $`N_i`$ becomes large. But this process of the $`N_i`$ changing roles takes a long time, and most of it occurs close to the Kasner circle, where $`\mathrm{\Omega }`$ decays exponentially. Thus, $`\mathrm{\Omega }`$ may increase by a certain factor, but after that it must decay by a larger factor until it can increase again, hence the result. Figure 6 illustrates the behaviour.
We divide the proof into lemmas, and begin by making the statement that $`\mathrm{\Omega }`$ decays exponentially close to the Kasner circle more precise.
###### Lemma 14.1.
Consider a Bianchi IX solution with $`2/3<\gamma <2`$. If
$$\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2\frac{1}{8}(3\gamma +2)$$
in an interval $`[s_1,s_2]`$, then
$$\mathrm{\Omega }(s)\mathrm{\Omega }(s_2)e^{\alpha _\gamma (s_2s)}$$
for $`s[s_1,s_2]`$, where
$$\alpha _\gamma =\frac{3}{2}(2\gamma ).$$
Proof. Observe that
(14.2)
$$\mathrm{\Omega }^{}4[\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2\frac{1}{4}(3\gamma 2)]\mathrm{\Omega },$$
so that under the conditions of the lemma
$$\mathrm{\Omega }^{}\alpha _\gamma \mathrm{\Omega }.$$
The conclusion follows. $`\mathrm{}`$
Next, we prove that if the $`N_i`$ all stay sufficiently small under a condition as in (14.1) and $`\mathrm{\Omega }`$ starts out small, then $`\mathrm{\Omega }`$ will remain small.
###### Lemma 14.2.
Consider a Bianchi IX solution with $`2/3<\gamma <2`$. There is an $`ฯต>0`$ such that if
(14.3) $`{\displaystyle \frac{3}{4}}N_i^2{\displaystyle \frac{1}{8}}(63\gamma )`$
(14.4) $`N_1N_2+N_2N_3+N_1N_3ฯต`$
in an interval $`[s_1,s_2]`$, and $`\mathrm{\Omega }(s_2)ฯต`$, then $`\mathrm{\Omega }(s)\mathrm{\Omega }(s_2)`$ for all $`s[s_1,s_2]`$.
Proof. Let
$$=\{\tau [s_1,s_2]:t[\tau ,s_2]\mathrm{\Omega }(t)\mathrm{\Omega }(s_2)\}.$$
Let $`\tau `$, $`\tau >s_1`$. There must be two $`N_i`$, say $`N_2`$ and $`N_3`$, such that $`N_2ฯต^{1/2}`$ and $`N_3ฯต^{1/2}`$ in $`\tau `$, by (14.4). By the constraint (2.3) and (14.4), we have in $`\tau `$,
$$\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^21\frac{3}{4}N_1^2\mathrm{\Omega }h_1\frac{1}{8}(3\gamma +2)4ฯต,$$
so that assuming $`ฯต`$ small enough depending only on $`\gamma `$, we have $`\mathrm{\Omega }^{}(\tau )>0`$, cf. (14.2). Thus there exists an $`s<\tau `$ such that $`s`$. In other words, $``$ is an open, closed, and non-empty subset of $`[s_1,s_2]`$, so that $`=[s_1,s_2]`$. $`\mathrm{}`$
The next lemma describes the phase during which $`\mathrm{\Omega }`$ may increase.
###### Lemma 14.3.
Consider a Bianchi IX solution with $`2/3<\gamma <2`$. There is an $`ฯต>0`$ such that if
(14.5) $`{\displaystyle \frac{3}{4}}N_1^2{\displaystyle \frac{1}{8}}(63\gamma )`$
(14.6) $`N_1N_2+N_2N_3+N_1N_3ฯต`$
in $`[s_1,s_2]`$, and $`\mathrm{\Omega }(s_2)ฯต`$, then $`s_2s_1c_{1,\gamma }`$ and $`\mathrm{\Omega }(s)c_{2,\gamma }\mathrm{\Omega }(s_2)`$ for all $`s[s_1,s_2]`$, where $`c_{1,\gamma }`$ and $`c_{2,\gamma }`$ are positive constants depending on $`\gamma `$.
Proof. Assume $`ฯต`$ is small enough that
$$\frac{3}{4}ฯต^{1/2}\frac{1}{8}(63\gamma ),$$
so that $`N_1ฯต^{1/4}`$ in $`[s_1,s_2]`$. Assuming $`ฯต<1`$ we get $`N_iฯต^{1/2}`$ in $`[s_1,s_2]`$, $`i=2,3`$. Use the constraint (2.3) to write
(14.7)
$$1\mathrm{\Omega }\mathrm{\Sigma }_+^2\mathrm{\Sigma }_{}^2=\frac{3}{4}N_1^2+h_1$$
where $`|h_1|3ฯต`$ by (14.6). Thus,
$$1\mathrm{\Omega }\mathrm{\Sigma }_+^2\mathrm{\Sigma }_{}^2\frac{3}{4}ฯต^{1/2}3ฯต,$$
so that we may assume
(14.8)
$$\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2<1$$
in $`[s_1,s_2]`$.
We now compare the behaviour with a type II vacuum solution. By (2.4) and (14.7), we have
(14.9)
$$\mathrm{\Sigma }_+^{}=2(\frac{3}{4}N_1^2+h_1)(\mathrm{\Sigma }_++1)\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_++\frac{9}{2}N_1^2$$
$$\frac{9}{2}N_1(N_2+N_3)=\frac{3}{2}N_1^2(2\mathrm{\Sigma }_+)+h_2\mathrm{\Omega }+h_3,$$
where $`|h_3|17ฯต`$ and $`|h_2|2`$ in $`[s_1,s_2]`$. Let $`a_\gamma =(63\gamma )/4`$. Then,
$$\mathrm{\Sigma }_+(s_2)\mathrm{\Sigma }_+(s_1)a_\gamma (s_2s_1)+_{s_1}^{s_2}(h_2\mathrm{\Omega }+h_3)๐t.$$
However,
$$\mathrm{\Omega }(s)\mathrm{\Omega }(s_2)e^{4(ss_2)}ฯตe^{4(ss_2)}$$
for all $`s[s_1,s_2]`$, see (2.1). Thus,
$$|_{s_1}^{s_2}h_2\mathrm{\Omega }๐s|\frac{1}{2}\mathrm{\Omega }(s_2)e^{4(s_2s_1)}.$$
We get
$$\mathrm{\Sigma }_+(s_2)\mathrm{\Sigma }_+(s_1)a_\gamma (s_2s_1)\frac{1}{2}ฯตe^{4(s_2s_1)}17ฯต(s_2s_1).$$
This inequality contradicts the statement that $`s_2s_1`$ may be taken equal to $`4/a_\gamma `$, by choosing $`ฯต`$ small enough. We conclude that $`s_2s_14/a_\gamma =c_{1,\gamma }`$, and that we may choose $`c_{2,\gamma }=\mathrm{exp}(16/a_\gamma )`$. $`\mathrm{}`$
The following lemma deals with the decay in $`\mathrm{\Omega }`$ that has to follow an increase. The idea is that if $`N_1`$ is on the boundary between big and small, and its derivative is non-negative at a point, then it will decrease as we go backward, and the solution will not move far from the Kasner circle until one of the other $`N_i`$ has become large. That takes a long time and $`\mathrm{\Omega }`$ will decay.
###### Lemma 14.4.
Consider a Bianchi IX solution such that $`2/3<\gamma <2`$. There is an $`ฯต>0`$ such that if
(14.10)
$$N_1N_2+N_2N_3+N_3N_1ฯต$$
in $`[s_1,s_2]`$,
$$\frac{3}{4}N_1^2(s_2)=\frac{1}{8}(63\gamma ),N_1^{}(s_2)0$$
and $`\mathrm{\Omega }(s_2)c_{2,\gamma }ฯต`$, where $`c_{2,\gamma }`$ is the constant appearing in Lemma 14.3, then $`\mathrm{\Omega }`$ decays as we go backward starting at $`s_2`$, until $`s=s_1`$, or we reach a point $`s`$ at which
$$\mathrm{\Omega }(s)\frac{\mathrm{\Omega }(s_2)}{2c_{2,\gamma }}.$$
Proof. We begin by assuming that $`ฯต>0`$ is a fixed number. As the proof progresses, we will restrict it to be smaller than a certain constant depending on $`\gamma `$. We could spell it out here, but prefer to add restrictions successively. Let $`N_1ฯต^{1/4}`$ in $`[t_1,s_2]`$ and $`N_1(t_1)=ฯต^{1/4}`$ or $`t_1=s_1`$, in case $`N_1`$ does not attain $`ฯต^{1/4}`$ in $`[s_1,s_2]`$. As in the proof of Lemma 14.3, we conclude that $`N_iฯต^{1/2}`$, $`i=2,3`$ in $`[t_1,s_2]`$, and that we may assume
(14.11)
$$\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2<1.$$
The variables $`(\mathrm{\Omega },\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ have to belong to the interior of a paraboloid for $`N_1^{}`$ to be negative. Since $`N_1^{}(s_2)0`$ we are on the boundary or outside the paraboloid. The boundary is given by $`g=0`$, where
$$g=\frac{1}{2}(3\gamma 2)\mathrm{\Omega }+2\mathrm{\Sigma }_+^2+2\mathrm{\Sigma }_{}^24\mathrm{\Sigma }_+.$$
An outward pointing normal is given by $`g`$, where the derivatives are taken in the order: $`\mathrm{\Omega }`$, $`\mathrm{\Sigma }_+`$ and $`\mathrm{\Sigma }_{}`$. Let
$$=\{\tau [t_1,s_2]:t[\tau ,s_2]N_1^{}(t)0,\mathrm{\Omega }(t)c_{2,\gamma }ฯต\}.$$
Let $`\tau `$. By (14.11) we get $`q(\tau )<2`$ and, as we are also outside the interior of the paraboloid, $`\mathrm{\Sigma }_+(\tau )1/2`$. For $`ฯต`$, and thereby $`\mathrm{\Omega }`$, small enough depending only on $`\gamma `$, we have
$$\mathrm{\Sigma }_+^{}(\tau )ฯต^{1/2},$$
cf. (14.9). Using the above observations, we estimate in $`\tau `$,
$$g(\mathrm{\Omega }^{},\mathrm{\Sigma }_+^{},\mathrm{\Sigma }_{}^{})C_\gamma ฯตฯต^{1/2},$$
where $`C_\gamma `$ only depends on $`\gamma `$. For $`ฯต`$ small enough, the scalar product is negative. Thus, if $`(\mathrm{\Omega }(\tau ),\mathrm{\Sigma }_+(\tau ),\mathrm{\Sigma }_{}(\tau ))`$ is on the surface of the paraboloid, the solution moves away from it as we go backward, so that $`N_1^{}0`$ in $`[s,\tau ]`$ for some $`s<\tau `$. If we are already outside the paraboloid, the existence of such an $`s`$ is guaranteed by less complicated arguments. As in the proof of Lemma 14.2, we get $`\mathrm{\Omega }^{}>0`$ for $`ฯต`$ small enough depending only on $`\gamma `$, so that $``$ is open, closed and non-empty. Thus $`N_1`$ decreases from $`s_2`$ to $`t_1`$ going backward. Now,
$$\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^21\frac{3}{4}N_1^2\mathrm{\Omega }h_1\frac{1}{8}(3\gamma +2)c_{2,\gamma }ฯต3ฯต$$
in $`[t_1,s_2]`$, so that
(14.12)
$$\mathrm{\Omega }(t_1)\mathrm{\Omega }(s_2)e^{(2\gamma )(s_2t_1)},$$
by an argument similar to Lemma 14.1, if $`ฯต`$ is small enough. We can assume $`ฯต`$ is small enough that the time required for $`N_1`$ to decrease to $`ฯต^{1/4}`$ is great enough that if $`t_1s_1`$, then the conclusion of the lemma follows by (14.12). $`\mathrm{}`$
Proof of Proposition 14.1. Assume $`ฯต`$ is small enough that all the conditions of Lemma 14.2-14.4 are fulfilled. We divide the interval $`[\tau _1,\tau _2]`$ into suitable subintervals, such that we may apply the above lemmas to them. If
(14.13)
$$\frac{3}{4}N_i^2\frac{1}{8}(63\gamma )$$
in $`\tau _2`$ for $`i=1,2,3`$, then we let $`t_2[\tau _1,\tau _2]`$ be the smallest member of the interval such that (14.13) holds in all of $`[t_2,\tau _2]`$. Otherwise, we chose $`t_2=\tau _2`$. Either $`t_2=\tau _1`$ or $`3N_1^2(t_2)/4(63\gamma )/8`$, by a suitable permutation of the variables. If $`t_2\tau _1`$, let $`t_1`$ be the smallest member of $`[\tau _1,t_2]`$ such that $`3N_1^2/4(63\gamma )/8`$ in $`[t_1,t_2]`$.
Because of Lemma 14.2, $`\mathrm{\Omega }`$ decays in $`[t_2,\tau _2]`$. If $`t_2=\tau _1`$, we are done; let $`c_\gamma =1`$. Otherwise, we apply Lemma 14.3 to the interval $`[t_1,t_2]`$ to conclude that $`\mathrm{\Omega }(\tau )c_{2,\gamma }\mathrm{\Omega }(\tau _2)`$ in $`[t_1,\tau _2]`$. If $`t_1=\tau _1`$, we can choose $`c_\gamma =c_{2,\gamma }`$. Otherwise, we apply Lemma 14.4 to $`[\tau _1,t_1]`$. Either $`\mathrm{\Omega }`$ decays until we have reached $`\tau _1`$, or there is a point $`s_1[\tau _1,t_1]`$ such that $`\mathrm{\Omega }(s_1)\mathrm{\Omega }(\tau _2)/2`$. By the proof of Lemma 14.4, we can assume that $`\tau _2s_11`$; some time has to elapse for the decay to take place.
Given an interval $`[\tau _1,\tau _2]`$ as in the statement of the proposition, there are thus two possibilities. Either $`\mathrm{\Omega }(\tau )c_{2,\gamma }\mathrm{\Omega }(\tau _2)`$ for all $`\tau [\tau _1,\tau _2]`$ or we can construct an $`s_1[\tau _1,\tau _2]`$ such that $`\tau _2s_11`$, $`\mathrm{\Omega }(s_1)\mathrm{\Omega }(\tau _2)/2`$, and $`\mathrm{\Omega }(\tau )c_{2,\gamma }\mathrm{\Omega }(\tau _2)`$ for all $`\tau [s_1,\tau _2]`$. If the second possibility is the one that occurs, we can apply the same argument to $`[\tau _1,s_1]`$, and by repeated application, the proposition follows. $`\mathrm{}`$
###### Corollary 14.1.
Consider a Bianchi IX solution with $`2/3<\gamma <2`$. If
$$\underset{\tau \mathrm{}}{lim}(N_1N_2+N_2N_3+N_1N_3)=0$$
and there is a sequence $`\tau _k\mathrm{}`$ such that $`\mathrm{\Omega }(\tau _k)0`$, then
$$\underset{\tau \mathrm{}}{lim}\mathrm{\Omega }(\tau )=0.$$
## 15. Generic attractor for Bianchi IX solutions
In this section, we prove that for a generic Bianchi IX solution, the closure of the set of type II vacuum points is an attractor, assuming $`2/3<\gamma <2`$. What we need to prove is that
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Omega }+N_1N_2+N_2N_3+N_1N_3)=0,$$
since then we may for each $`ฯต>0`$ choose a $`T`$ such that at least two of the $`N_i`$ and $`\mathrm{\Omega }`$ must be less than $`ฯต`$ for $`\tau T`$. The starting point is the existence of a limit point on the Kasner circle for a generic solution, given by Proposition 13.1. Since there is such a limit point, there is a sequence $`\tau _k\mathrm{}`$ such that $`N_i(\tau _k)`$ and $`\mathrm{\Omega }(\tau _k)`$ go to zero. If
(15.1)
$$h=N_1N_2+N_2N_3+N_1N_3$$
does not converge to zero, it must thus grow from an arbitrarily small value up to some $`ฯต`$. By choosing $`ฯต`$ so that Proposition 14.1 is applicable, we have control over $`\mathrm{\Omega }`$. A few arguments yield the conclusion that we may assume that it is the product $`N_2N_3`$ that grows, and that the growth occurs close to the special point $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(1,0)`$. Close to this point, $`\mathrm{\Omega }`$, $`N_1`$ and $`N_1(N_2+N_3)`$ decay exponentially, so as far as intuition goes, we may equate them with zero. We thus have a Bianchi VI$`\mathrm{I}_0`$ vacuum solution close to the special point $`(1,0)`$. The behaviour of $`N_2N_3`$ will be oscillatory, and we may reduce the problem to one in which the product behaves essentially as a sine wave. However, by doing some technical estimates, one may see that one goes down going from top to top during the oscillation, and that that contradicts the assumed growth. Figure 7 illustrates the behaviour. It is a simulation of part of a Bianchi VI$`\mathrm{I}_0`$ vacuum solution.
We begin by rewriting the solutions in a form that makes the oscillatory behaviour apparent. Consider a non Taub-NUT Bianchi IX solution in an interval such that $`1<\mathrm{\Sigma }_+<1`$. Define the functions
(15.2) $`\stackrel{~}{x}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Sigma }_{}}{(1\mathrm{\Sigma }_+^2)^{1/2}}}`$
(15.3) $`\stackrel{~}{y}`$ $`=`$ $`{\displaystyle \frac{\sqrt{3}}{2}}{\displaystyle \frac{N_2N_3}{(1\mathrm{\Sigma }_+^2)^{1/2}}}.`$
The reason why these expressions are natural to consider is that, for reasons mentioned above, $`N_1`$, $`\mathrm{\Omega }`$ and so forth may be considered to be zero. In the situation we will need to consider $`N_2N_3`$ and $`\mathrm{\Sigma }_{}`$ will have much greater derivatives than $`\mathrm{\Sigma }_+`$, so that it is natural to consider $`\stackrel{~}{x}`$ and $`\stackrel{~}{y}`$ as sine and cosine, since the constraint essentially says $`\stackrel{~}{x}^2+\stackrel{~}{y}^2=1`$. Let
(15.4)
$$g=3(N_2+N_3)2(1+\mathrm{\Sigma }_+)\stackrel{~}{x}\stackrel{~}{y}=g_1+g_2.$$
In our applications, $`g_1`$ will essentially be constant, and $`g_2`$ will essentially be zero.
###### Lemma 15.1.
The vector $`\stackrel{~}{๐ฑ}=(\stackrel{~}{x},\stackrel{~}{y})^t`$ satisfies
$$\stackrel{~}{๐ฑ}^{}=A\stackrel{~}{๐ฑ}+ฯต,$$
where $`A`$ is defined as in (12.1), with $`g`$ as in (15.4) and $`ฯต=(ฯต_x,ฯต_y)^t`$, where the components are given by (15.5) and (15.6).
The error terms are
(15.5)
$$ฯต_x=3N_1\stackrel{~}{y}+(\frac{9}{2}N_1(N_1N_2N_3)\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_+)\frac{\mathrm{\Sigma }_+\stackrel{~}{x}}{1\mathrm{\Sigma }_+^2}$$
$$(\frac{3}{2}N_1^23N_1(N_2+N_3))\frac{\mathrm{\Sigma }_+\stackrel{~}{x}}{1\mathrm{\Sigma }_+}\frac{3}{2}(2\gamma )\mathrm{\Omega }\stackrel{~}{x}2(\frac{3}{4}N_1^2\frac{3}{2}N_1(N_2+N_3))\stackrel{~}{x}$$
and
(15.6)
$$ฯต_y=[\frac{1}{2}(3\gamma 2)\mathrm{\Omega }(1+\mathrm{\Sigma }_+)+\frac{3}{2}(2\gamma )\mathrm{\Omega }+\frac{9}{2}N_1(N_1N_2N_3)]\frac{\stackrel{~}{y}\mathrm{\Sigma }_+}{1\mathrm{\Sigma }_+^2}$$
$$+\frac{1}{2}(3\gamma 2)\mathrm{\Omega }\stackrel{~}{y}.$$
It is clear that if we have a vacuum type VI$`\mathrm{I}_0`$ solution, $`ฯต_x=ฯต_y=0`$, so that we may write $`\stackrel{~}{๐ฑ}=(\mathrm{sin}(\xi (\tau )),\mathrm{cos}(\xi (\tau )))`$, where $`\xi `$ is as in (12.2). In our situation, there is an error term, but by the exponential decay mentioned above, it only makes the technical details somewhat longer.
We begin by proving that we can assume that the growth occurs in the product $`N_2N_3`$, and that $`\mathrm{\Omega }`$ can be assumed to be negligible during the growth. We also put bounds on $`\mathrm{\Sigma }_+`$. They constitute a starting point for further restrictions. The values of certain constants have been chosen for future convenience.
The lemma below is formulated to handle more general situations than the one above. One reason being the desire to prove uniform convergence to the attractor. We will use the terminology that if $`x`$ constitutes initial data for (2.1)-(2.3), then $`\mathrm{\Sigma }_+(\tau ,x)`$ and so on will denote the solution of the equations with initial value $`x`$ evaluated at $`\tau `$, assuming that $`\tau `$ belongs to the existence interval. We will use $`\mathrm{\Phi }(\tau ,x)`$ to summarize all the variables. The goal of this section is to prove that the conditions of the lemma below are never met.
###### Lemma 15.2.
Let $`2/3<\gamma <2`$. Consider a sequence $`x_l`$ of Bianchi IX initial data with all $`N_i>0`$ and two sequences $`s_l\tau _l`$ of real numbers, belonging to the existence interval corresponding to $`x_l`$, such that
(15.7)
$$\underset{l\mathrm{}}{lim}d(\tau _l,x_l)=0,$$
where $`d=\mathrm{\Omega }+N_1N_2+N_2N_3+N_1N_3`$, and
(15.8)
$$h(s_l,x_l)\delta $$
for some $`\delta >0`$ independent of $`l`$. Then there is an $`ฯต>0`$ and a $`k_0`$, such that for each $`kk_0`$ there is an $`l_k`$, a symmetry operation on $`\mathrm{\Phi }(,x_{l_k})`$, and an interval $`[u_k,v_k]`$ belonging to the existence interval of $`\mathrm{\Phi }(,x_{l_k})`$, such that the transformed variables satisfy
$$(N_2N_3)(u_k,x_{l_k})=ฯต,(N_2N_3)(v_k,x_{l_k})ฯตe^{20k},ฯตe^{20k1}(N_2N_3)(\tau ,x_{l_k})ฯต$$
(15.9)
$$N_1(\tau ,x_{l_k})ฯต\mathrm{exp}(30k)and2N_2(\tau ,x_{l_k}),N_3(\tau ,x_{l_k})ฯต\mathrm{exp}(25k)$$
for $`\tau [u_k,v_k]`$. Furthermore
(15.10)
$$\mathrm{\Omega }(,x_{l_k})e^{13k}\mathrm{and}1<\mathrm{\Sigma }_+(,x_{l_k})0$$
in $`[u_k,v_k]`$.
Remark. Observe that for the main application of this lemma, the sequence $`x_l`$ will be independent of $`l`$.
Proof. By (15.7) and (15.8), there is an $`ฯต>0`$ such that for every $`k`$ there is a suitable $`l_k`$ and $`u_kv_k`$ with $`[u_k,v_k][s_{l_k},\tau _{l_k}]`$ such that
(15.11)
$$e^{20k1}ฯตh(\tau ,x_{l_k})2ฯต$$
$`h(u_k,x_{l_k})=2ฯต`$, $`h(v_k,x_{l_k})=\mathrm{exp}(20k1)ฯต`$ where $`\tau [u_k,v_k]`$. We can also assume that
(15.12)
$$h(\tau ,x_{l_k})2ฯต$$
for all $`\tau [u_k,\tau _{l_k}]`$. Furthermore, we can assume
(15.13)
$$(N_1N_2N_3)(,x_{l_k})ฯต^2\mathrm{exp}(50k1)/4$$
in $`[u_k,\tau _{l_k}]`$. The reason is that $`d(\tau _l,x_l)`$ converges to zero, so that $`(N_1N_2N_3)(\tau _l,x_l)`$ also converges to zero. Consequently, we can assume $`(N_1N_2N_3)(\tau _{l_k},x_{l_k})`$ to be as small as we wish, and thus we get (15.13) by the monotonicity of the product. Since we may assume $`\mathrm{\Omega }(\tau _{l_k},x_{l_k})`$ to be arbitrarily small by (15.7), we may apply Proposition 14.1 in $`[u_k,\tau _{l_k}]`$ by (15.12), choosing $`ฯต`$ small enough. Thus we may assume $`\mathrm{\Omega }\mathrm{exp}(13k)`$ in $`[u_k,v_k]`$. From now on, we consider the solution $`\mathrm{\Phi }(,x_{l_k})`$ in the interval $`[u_k,\tau _{l_k}]`$ and only use the observations above. To avoid cumbersome notation, we will omit reference to the evaluation at $`x_{l_k}`$. By (15.11) and (15.13), we have in $`[u_k,v_k]`$
$$ฯตe^{20k1}h=N_1N_2N_3(\frac{1}{N_1}+\frac{1}{N_2}+\frac{1}{N_3})\frac{1}{4}ฯต^2e^{50k1}(\frac{1}{N_1}+\frac{1}{N_2}+\frac{1}{N_3}),$$
so that
$$\frac{1}{N_1}+\frac{1}{N_2}+\frac{1}{N_3}\frac{4}{ฯต}e^{30k}.$$
At a given $`\tau [u_k,v_k]`$, one $`N_i`$, say $`N_1`$, must be smaller than $`ฯต\mathrm{exp}(30k)`$. If the second smallest is smaller than $`ฯต\mathrm{exp}(25k)`$, the largest cannot be bigger than $`2`$, by Lemma 3.3, but that will contradict $`hฯต\mathrm{exp}(20k1)`$ if $`k`$ is great enough. Thus, if $`N_1`$ is the smallest $`N_i`$ for one $`\tau `$, it is always the smallest. We may thus assume
$$N_1ฯต\mathrm{exp}(30k)\mathrm{and}N_2,N_3ฯต\mathrm{exp}(25k)$$
in $`[u_k,v_k]`$. If $`ฯต`$ is small enough, we can assume $`N_2,N_32`$ by Lemma 3.3. Thus,
$$e^{20k1}ฯต4ฯตe^{30k}N_2N_32ฯต+4ฯตe^{30k}.$$
We may shift $`u_k`$ by adding a positive number to it so that
(15.14)
$$(N_2N_3)(u_k)=ฯต\mathrm{and}(N_2N_3)(\tau )ฯต$$
for $`\tau [u_k,v_k]`$. We may also shift $`v_k`$ in the negative direction to achieve
$$(N_2N_3)(v_k)ฯตe^{20k},(N_2N_3)^{}(v_k)<0\mathrm{and}(N_2N_3)(\tau )ฯตe^{20k1}$$
for $`\tau [u_k,v_k]`$. The condition on the derivative is there to get control on $`\mathrm{\Sigma }_+`$.
We now establish rough control of $`\mathrm{\Sigma }_+`$. Since $`(N_2N_3)^{}(v_k)<0`$, $`1<\mathrm{\Sigma }_+(v_k)<0`$. Due to (15.9), (2.4) and the constraint, $`\mathrm{\Sigma }_+^{}<0`$ if $`\mathrm{\Sigma }_+=0`$ or $`\mathrm{\Sigma }_+=1`$. In other words, $`\mathrm{\Sigma }_+(w_k)=0`$ implies $`\mathrm{\Sigma }_+0`$ in $`[u_k,w_k]`$. But if $`u_k<w_k`$ then $`\mathrm{\Sigma }_+(u_k)>0`$ so that $`(N_2N_3)(u_k)<(N_2N_3)(w_k)`$, contradicting the construction as stated in (15.14). We thus have $`\mathrm{\Sigma }_+0`$ in $`[u_k,v_k]`$. We also have $`1<\mathrm{\Sigma }_+`$ in that interval. $`\mathrm{}`$
Below, we will omit reference to the evaluation at $`x_{l_k}`$ to avoid cumbersome notation, but it should be remembered that we in general have a different solution for each $`k`$. Let
$$r(\tau )=_\tau ^{v_k}(q/2+\mathrm{\Sigma }_+)๐s.$$
Here we mean $`q(s,x_{l_k})`$ when we write $`q`$, and similarly for $`\mathrm{\Sigma }_+`$. Observe that $`r`$ depends on $`k`$, but that we omit reference to this dependence. All the information concerning the growth of $`N_2N_3`$ is contained in $`r`$, see (2.1), and this integral will be our main object of study rather than the product $`N_2N_3`$. Let $`[u_k,v_k]`$ be an interval as in Lemma 15.2. Since
$$(N_2N_3)(v_k)=e^{4r(u_k)}(N_2N_3)(u_k),$$
we have $`r(u_k)5k`$. Let $`u_k\nu _k\sigma _k\tau _kr_kv_k`$. Starting at $`u_k`$, let $`\nu _k`$ be the last point $`r=4k`$, so that $`r4k`$ in $`[\nu _k,v_k]`$. Furthermore, let $`rk`$ in $`[r_k,v_k]`$ and finally, assume $`r2k`$ in $`[\nu _k,\tau _k]`$. We also assume that $`r`$ evaluated at $`r_k`$, $`\tau _k`$, $`\sigma _k`$ and $`\nu _k`$ is $`k`$, $`2k`$, $`3k`$ and $`4k`$ respectively. See Table 2. Why? The interval we will work with in the end is $`[\sigma _k,\tau _k]`$, but the other intervals are used to get control of the variables there. First of all, we want to get control of $`\mathrm{\Sigma }_+`$, and the interval $`[u_k,\nu _k]`$ together with the additional demand on $`\nu _k`$ serves that purpose. The intervals at the other end, together with the associated demands, are there to yield us a quantitative statement of the intuitive idea that $`\mathrm{\Omega }`$ and $`N_1`$ are negligible relative to the other expressions of interest. Finally, we need to get quantitative bounds relating the different variables; as was mentioned earlier, the main idea is to prove that $`N_2N_3`$ oscillates, but that it decreases during a period. In order to prove the decrease, we need to have control over the relative sizes of different expressions, and $`[\nu _k,\sigma _k]`$ is used to achieve the desired estimates.
From this point until the statement of Theorem 15.1, we will assume that the conditions of Lemma 15.2 are fulfilled. We will use the consequences of this assumption, as stated above, freely.
We improve the control of $`\mathrm{\Sigma }_+`$. Let us first give an intuitive argument. Observe that under the present circumstances, the solution is approximated by a Bianchi VI$`\mathrm{I}_0`$ vacuum solution. For such a solution, the function $`Z_1`$, defined in (13.2), is monotone increasing going backwards. According to the Bianchi VI$`\mathrm{I}_0`$ vacuum constraint, $`Z_1`$ is proportional to $`(1\mathrm{\Sigma }_+^2)/N_2N_3`$. However, we know that $`N_2N_3`$ has to increase by a factor of $`e^{20k}`$ going from $`v_k`$ to $`u_k`$, and consequently $`1\mathrm{\Sigma }_+^2`$ has to increase by an even larger factor. The only way this can occur, is if a large part of the growth in $`N_2N_3`$ occurs when $`\mathrm{\Sigma }_+`$ is very close to $`1`$. Taking this into account, we see that the relevant variation in $`1\mathrm{\Sigma }_+^2=(1\mathrm{\Sigma }_+)(1+\mathrm{\Sigma }_+)`$ occurs in the factor $`1+\mathrm{\Sigma }_+`$. Below, we will use the function $`(1+\mathrm{\Sigma }_+)/N_2N_3`$ instead of $`Z_1`$. Let us begin by considering the vacuum case, in order to see the idea behind the argument, without the technical difficulties associated with the non-vacuum case. We have
(15.15)
$$\left(\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\right)^{}<0$$
in our situation, cf. Lemma 15.3 and (15.10). For $`\tau [\nu _k,v_k]`$ we get
$$0<1+\mathrm{\Sigma }_+(\tau )(1+\mathrm{\Sigma }_+(u_k))\frac{(N_2N_3)(\tau )}{(N_2N_3)(u_k)}e^{4k}$$
by our construction.
Let us make some observations before we turn to the non-vacuum case. First we analyze the derivative of $`(1+\mathrm{\Sigma }_+)/N_2N_3`$ in general. The estimates (15.16) and (15.17) will in fact be important throughout this section.
###### Lemma 15.3.
Let $`u_k`$ and $`v_k`$ be as above. Then
(15.16)
$$\left(\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\right)^{}\frac{2[(1+\mathrm{\Sigma }_+)^2+\mathrm{\Sigma }_{}^2](1+\mathrm{\Sigma }_+)+\frac{3}{2}(2\gamma )\mathrm{\Omega }}{N_2N_3}$$
and
(15.17)
$$\mathrm{\Sigma }_+^{}\frac{3}{2}(2\gamma )\mathrm{\Omega }$$
in the interval $`[u_k,v_k]`$ for $`k`$ large enough.
Remark. Observe that $`1+\mathrm{\Sigma }_+>0`$ in $`[u_k,v_k]`$ by (15.10), so that the first term appearing in the numerator of the right hand side of (15.16) has the right sign.
Proof. Using (2.4), we have
$$\left(\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\right)^{}=[(22\mathrm{\Omega }2\mathrm{\Sigma }_+^22\mathrm{\Sigma }_{}^2)(\mathrm{\Sigma }_++1)\frac{3}{2}(2\gamma )\mathrm{\Omega }\mathrm{\Sigma }_++$$
$$+\frac{9}{2}N_1(N_1N_2N_3)(2q+4\mathrm{\Sigma }_+)(1+\mathrm{\Sigma }_+)](N_2N_3)^1.$$
Consider the numerator of the right hand side. The term involving the $`N_i`$ has the right sign by (15.9), and the terms not involving $`\mathrm{\Omega }`$ add up to the first term of the numerator of the right hand side of (15.16). Let us consider the terms involving $`\mathrm{\Omega }`$. They are
$$2\mathrm{\Omega }(1+\mathrm{\Sigma }_+)\frac{3}{2}(2\gamma )\mathrm{\Omega }(1+\mathrm{\Sigma }_+)+\frac{3}{2}(2\gamma )\mathrm{\Omega }(3\gamma 2)\mathrm{\Omega }(1+\mathrm{\Sigma }_+)=$$
$$=\frac{1}{2}(3\gamma 2)\mathrm{\Omega }(1+\mathrm{\Sigma }_+)+\frac{3}{2}(2\gamma )\mathrm{\Omega }\frac{3}{2}(2\gamma )\mathrm{\Omega }$$
proving (15.16). To prove (15.17), we observe that by the constraint and the fact that $`0<1+\mathrm{\Sigma }_+1`$ in the interval of interest, we have
$$(22\mathrm{\Omega }2\mathrm{\Sigma }_+^22\mathrm{\Sigma }_{}^2)(\mathrm{\Sigma }_++1)3N_1(N_2+N_3)(1+\mathrm{\Sigma }_+)3N_1(N_2+N_3).$$
Inserting this inequality into (2.4), we get
$$\mathrm{\Sigma }_+^{}\frac{3}{2}(2\gamma )\mathrm{\Omega }(1+\mathrm{\Sigma }_+)+\frac{3}{2}(2\gamma )\mathrm{\Omega }+\frac{1}{2}N_1(9N_13N_23N_3)\frac{3}{2}(2\gamma )\mathrm{\Omega }$$
by (15.9) and (15.10) if $`k`$ is large enough, proving (15.17). $`\mathrm{}`$
In the vacuum case, $`\mathrm{\Sigma }_+`$ is monotone in our situation, see (15.17), but in the general case we have the following weaker result.
###### Lemma 15.4.
Consider an interval $`[s,t][u_k,v_k]`$ such that
$$\mathrm{\Sigma }_+^2\frac{1}{8}(3\gamma +2).$$
Then
(15.18)
$$(1+\mathrm{\Sigma }_+(t))\mathrm{\Omega }(t)1+\mathrm{\Sigma }_+(s)$$
if $`k`$ is large enough.
Proof. In $`[s,t]`$ we have
$$\mathrm{\Omega }^{}\alpha _\gamma \mathrm{\Omega },$$
where $`\alpha _\gamma =3(2\gamma )/2`$, see the proof of Lemma 14.1. Thus,
$$\mathrm{\Omega }(u)\mathrm{\Omega }(t)\mathrm{exp}[\alpha _\gamma (ut)]$$
for all $`u[s,t]`$. Integrating (15.17) we get (15.18). $`\mathrm{}`$
In connection with (15.16), the following lemma is of interest.
###### Lemma 15.5.
If $`k`$ is large enough and
$$(1+\mathrm{\Sigma }_+(\tau ))^3e^{3k}\mathrm{\Omega }(\tau )$$
for some $`\tau [u_k,v_k]`$, then
$$(1+\mathrm{\Sigma }_+)^3\frac{3}{4}(2\gamma )\mathrm{\Omega }$$
in $`[u_k,\tau ]`$.
Proof. If the solution is of vacuum type the lemma follows, so assume $`\mathrm{\Omega }>0`$. Let us first prove that $`(1+\mathrm{\Sigma }_+(u))^3e^k\mathrm{\Omega }(\tau )`$ for $`u[u_k,\tau ]`$. Assume there is an $`s[u_k,\tau ]`$ such that the reverse inequality holds. Then there is a $`t`$ with $`\tau ts`$, such that $`(1+\mathrm{\Sigma }_+)^3e^{3k}\mathrm{\Omega }(\tau )`$ in $`[s,t]`$, with equality at $`t`$. Because of (15.10), Lemma 15.4 is applicable for $`k`$ large enough. Thus
(15.19)
$$e^k\mathrm{\Omega }^{1/3}(\tau )\mathrm{\Omega }(t)1+\mathrm{\Sigma }_+(s)e^{k/3}\mathrm{\Omega }^{1/3}(\tau ).$$
However, by the proof of Lemma 15.2, Proposition 14.1 is applicable in any subinterval of $`[u_k,v_k]`$, so that $`\mathrm{\Omega }(t)c_\gamma \mathrm{\Omega }(\tau )`$. Substituting this into (15.19), we get
$$e^k\mathrm{\Omega }^{1/3}(\tau )c_\gamma \mathrm{\Omega }(\tau )e^{k/3}\mathrm{\Omega }^{1/3}(\tau ),$$
which is impossible for $`k`$ large enough.
Thus we have, for $`u[u_k,\tau ]`$ and $`k`$ large enough,
$$(1+\mathrm{\Sigma }_+(u))^3e^k\mathrm{\Omega }(\tau )e^k\frac{1}{\frac{3}{4}(2\gamma )c_\gamma }\frac{3}{4}(2\gamma )\mathrm{\Omega }(u)\frac{3}{4}(2\gamma )\mathrm{\Omega }(u)$$
where $`c_\gamma `$ is the constant appearing in the statement of Proposition 14.1. The lemma follows. $`\mathrm{}`$
We now prove that we have control over $`1+\mathrm{\Sigma }_+`$ in $`[\nu _k,v_k]`$.
###### Lemma 15.6.
Let $`\nu _k`$ and $`v_k`$ be as above. Then for $`k`$ large enough,
(15.20)
$$0<1+\mathrm{\Sigma }_+<e^k$$
in $`[\nu _k,v_k]`$.
Proof. Assume $`1+\mathrm{\Sigma }_+(\tau )e^k`$ for some $`\tau [\nu _k,v_k]`$. Because of (15.10), we then conclude that Lemma 15.5 is applicable, so that
$$\left(\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\right)^{}0$$
in $`[u_k,\tau ]`$ by (15.16). Thus
$$\frac{1+\mathrm{\Sigma }_+(u_k)}{(N_2N_3)(u_k)}\frac{1+\mathrm{\Sigma }_+(\tau )}{(N_2N_3)(\tau )}\frac{e^k}{(N_2N_3)(\tau )},$$
but by our construction
$$(N_2N_3)(\tau )=e^{4r(u_k)4r(\tau )}(N_2N_3)(u_k)e^{20k+16k}(N_2N_3)(u_k),$$
so that
$$e^{3k}1+\mathrm{\Sigma }_+(u_k)1.$$
The lemma follows. $`\mathrm{}`$
###### Corollary 15.1.
Let $`\nu _k`$ and $`v_k`$ be as above. For $`k`$ large enough,
$$\mathrm{\Omega }+\mathrm{\Sigma }_{}^2+(1+\mathrm{\Sigma }_+)^24e^k$$
in $`[\nu _k,v_k]`$.
Proof. By (15.9), we have
$$N_1(N_2+N_3)4ฯตe^{30k}$$
in $`[u_k,v_k]`$. This observation, the constraint, and Lemma 15.6 yield
$$\mathrm{\Omega }+\mathrm{\Sigma }_{}^21\mathrm{\Sigma }_+^2+\frac{3}{2}N_1(N_2+N_3)3e^k$$
in $`[\nu _k,v_k]`$, for $`k`$ large enough. The corollary follows using Lemma 15.6. $`\mathrm{}`$
The next thing to prove is that $`N_1`$ and $`\mathrm{\Omega }`$ are small compared with $`1+\mathrm{\Sigma }_+`$. The fact that $`r(r_k)=k`$ will imply that the integral of $`1+\mathrm{\Sigma }_+`$ is large, but if $`1+\mathrm{\Sigma }_+`$ is comparable with $`N_1`$ or $`\mathrm{\Omega }`$, it cannot be large since $`N_1`$ and $`\mathrm{\Omega }`$ decay exponentially.
The reason $`(1+\mathrm{\Sigma }_+)^9`$ appears in the estimate (15.21) below is that the final argument will consist of an estimate of an integral up to โorder of magnitudeโ. Expressions of the form $`(1+\mathrm{\Sigma }_+)^n`$ and $`(1+\mathrm{\Sigma }_+)^m/(N_2+N_3)^l`$ will will define what is โbigโ and โsmallโ, and here we see to it that terms involving $`\mathrm{\Omega }`$ and $`N_1`$ are negligible in this order of magnitude calculus. Finally, the factor $`\mathrm{exp}(3k)`$ is there in order for us to be able to ignore possible factors multiplying expressions involving $`N_1`$ and $`\mathrm{\Omega }`$. We only turn up the number $`k`$ and change $`\mathrm{exp}(3k)`$ to $`\mathrm{exp}(2k)`$ to eliminate constants we do not want to think about; consider (15.5) and (15.6).
###### Lemma 15.7.
Let $`\nu _k`$ and $`\tau _k`$ be as above. Then for $`k`$ large enough,
(15.21)
$$\mathrm{\Omega }+N_1+N_1(N_2+N_3)e^{3k}e^{3b_\gamma (\tau v_k)}(1+\mathrm{\Sigma }_+)^9$$
in $`[\nu _k,\tau _k]`$ where $`b_\gamma >0`$. Furthermore,
(15.22)
$$\left(\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\right)^{}2\mathrm{\Sigma }_{}^2\frac{(1+\mathrm{\Sigma }_+)}{N_2N_3}$$
in $`[u_k,\tau _k]`$.
Proof. Note that
$$_{r_k}^{v_k}(1+\mathrm{\Sigma }_+)๐\tau _{r_k}^{v_k}(\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_+)๐\tau _{r_k}^{v_k}(q/2+\mathrm{\Sigma }_+)๐\tau =k,$$
so that
(15.23)
$$k_{r_k}^{v_k}(1+\mathrm{\Sigma }_+)๐\tau .$$
Let
$$\rho _1=\mathrm{\Omega }+N_1+N_1(N_2+N_3).$$
By the construction in Lemma 15.2, we may assume
$$\rho _1(v_k)e^{12k}.$$
Because of Corollary 15.1, we have
$$\rho _1(\tau )e^{12k}e^{4b_\gamma (\tau v_k)}$$
for all $`\tau [\nu _k,v_k]`$, where $`b_\gamma >0`$ is some constant depending only on $`\gamma `$. Let
$$\rho _2(\tau )=e^{9k}e^{b_\gamma (\tau v_k)}e^{3k}e^{3b_\gamma (\tau v_k)}\rho _1(\tau ).$$
The assumption that $`(1+\mathrm{\Sigma }_+)^9\rho _2`$ in $`[r_k,v_k]`$ contradicts (15.23). Thus there must be a $`t_0[r_k,v_k]`$ such that $`(1+\mathrm{\Sigma }_+(t_0))^9\rho _2(t_0)`$. In the vacuum case, $`1+\mathrm{\Sigma }_+`$ increases as we go backward, and $`\rho _2`$ obviously decreases, and thus we are in that case able to conclude $`(1+\mathrm{\Sigma }_+)^9\rho _2`$ in $`[\nu _k,r_k]`$. In the general case, we observe that $`(1+\mathrm{\Sigma }_+(t_0))^3e^{3k}\mathrm{\Omega }(t_0)`$ by the above constructions. We get
$$\left(\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\right)^{}2\mathrm{\Sigma }_{}^2\frac{(1+\mathrm{\Sigma }_+)}{N_2N_3}$$
in $`[u_k,t_0]`$, by combining Lemma 15.5 and (15.16). Inequality (15.22) follows. Thus, if $`\tau [\nu _k,\tau _k]`$, we have
$$1+\mathrm{\Sigma }_+(\tau )\frac{(N_2N_3)(\tau )}{(N_2N_3)(t_0)}(1+\mathrm{\Sigma }_+(t_0))e^{4k}(1+\mathrm{\Sigma }_+(t_0)).$$
Consequently, we will have $`(1+\mathrm{\Sigma }_+(\tau ))^9\rho _2(\tau )`$, since $`1+\mathrm{\Sigma }_+`$ has increased from its value at $`t_0`$ and $`\rho _2`$ has decreased. The lemma follows. $`\mathrm{}`$
Next we establish a relation between $`1+\mathrm{\Sigma }_+`$ and the product $`N_2N_3`$. We prove that $`(1+\mathrm{\Sigma }_+)/(N_2N_3)`$ can be chosen arbitrarily small in the interval $`[\sigma _k,\tau _k]`$, by estimating it in $`\nu _k`$, and then comparing the integral of $`1+\mathrm{\Sigma }_+`$ from $`\nu _k`$ to $`\sigma _k`$ with the integral of $`\mathrm{\Sigma }_{}^2`$ over the same interval. The following lemma is the starting point.
###### Lemma 15.8.
Let $`\sigma _k,\tau _k`$ be as above. Then for $`k`$ large enough,
(15.24)
$$\frac{1+\mathrm{\Sigma }_+(\tau )}{(N_2N_3)(\tau )}\frac{1}{ฯต}\mathrm{exp}(2_{\nu _k}^{\sigma _k}\mathrm{\Sigma }_{}^2๐s)$$
if $`\tau [\sigma _k,\tau _k]`$. Furthermore,
$$\frac{1+\mathrm{\Sigma }_+(\tau )}{(N_2N_3)(\tau )}\frac{1}{ฯต}$$
in $`[u_k,\tau _k]`$.
Proof. The statement follows from (15.22), and the fact that
$$\frac{(1+\mathrm{\Sigma }_+)(u_k)}{(N_2N_3)(u_k)}\frac{1}{ฯต}.$$
$`\mathrm{}`$
Considering the constraint, it is clear that $`\mathrm{\Sigma }_{}^2`$ should be comparable with $`1+\mathrm{\Sigma }_+`$ when $`N_2N_3`$ and $`\mathrm{\Sigma }_{}`$ oscillate, and thus the integral should be comparable with $`k`$, cf. (15.23). However, we have to work out the technical details.
We carry out the comparison between the integrals in three steps. First, we estimate the error committed in viewing $`\stackrel{~}{x}`$ and $`\stackrel{~}{y}`$ in (15.2) and (15.3) as sine and cosine. Then we may, up to a small error, express the integral of $`\mathrm{\Sigma }_{}^2`$ as the integral of $`\mathrm{sin}^2(\eta /2)`$, multiplied by some function $`f(\eta )`$ by changing variables. In order to make the comparison, we need to estimate the variation of $`f`$ during a period: the second step. The only expressions involved are $`1+\mathrm{\Sigma }_+`$ and $`N_2+N_3`$. The third step consists of making the comparison, using the information obtained in the earlier steps.
Let $`\stackrel{~}{x}`$, $`\stackrel{~}{y}`$, $`g`$, $`g_1`$ and $`g_2`$ be defined as in (15.2)-(15.4), and $`\xi `$, $`x`$ and $`y`$ be defined as in the statement of Lemma 12.1, with $`\tau _0`$ replaced by $`\tau _k`$ and $`\varphi _0`$ by $`\varphi _k`$. Observe that $`x`$, $`y`$ and $`\xi `$ in fact depend on $`k`$. We need to compare $`x`$ with $`\stackrel{~}{x}`$.
###### Lemma 15.9.
Let $`\nu _k`$ and $`\tau _k`$ be as above. Then for $`k`$ large enough,
(15.25)
$$|\mathrm{\Sigma }_{}^2(1\mathrm{\Sigma }_+^2)x^2|12e^{2k}(1+\mathrm{\Sigma }_+)^9.$$
in $`[\nu _k,\tau _k]`$. Furthermore,
(15.26)
$$|1(\stackrel{~}{x}^2+\stackrel{~}{y}^2)|e^k$$
and
(15.27)
$$\stackrel{~}{๐ฑ}๐ฑ3e^{2k}(1+\mathrm{\Sigma }_+)^8$$
in that interval.
Proof. We have
$$|1(\stackrel{~}{x}^2(\tau _k)+\stackrel{~}{y}^2(\tau _k))^{1/2}||1\left(1+\frac{\frac{3}{2}N_1(N_2+N_3)\frac{3}{4}N_1^2\mathrm{\Omega }}{1\mathrm{\Sigma }_+^2}\right)^{1/2}|$$
(15.28)
$$e^{2k}(1+\mathrm{\Sigma }_+(\tau _k))^8$$
by (15.21). Equation (15.26) follows similarly. By (15.5), (15.6), (15.21) and (15.26), we have
$$ฯต(s)2b_\gamma e^{2k}(1+\mathrm{\Sigma }_+(s))^8e^{3b_\gamma (sv_k)}$$
for $`k`$ large enough. Let us estimate how much $`1+\mathrm{\Sigma }_+`$ may decrease as we go backward in time. By (15.17) and (15.21), we have
$$(1+\mathrm{\Sigma }_+)^{}\frac{3}{2}(2\gamma )e^{3k}e^{3b_\gamma (\tau v_k)}(1+\mathrm{\Sigma }_+)^9,$$
so that if $`[s,t][\nu _k,\tau _k]`$,
(15.29)
$$1+\mathrm{\Sigma }_+(t)\mathrm{exp}(\mathrm{exp}(2k))(1+\mathrm{\Sigma }_+(s)),$$
for $`k`$ large enough. Thus, for $`\tau \tau _k`$, we get
(15.30)
$$_\tau ^{\tau _k}ฯต(s)๐se^{2k}(1+\mathrm{\Sigma }_+(\tau ))^8.$$
By (12.4), (15.30), (15.29) and (15.28), we thus have
$$\stackrel{~}{๐ฑ}๐ฑ\frac{5}{2}e^{2k}(1+\mathrm{\Sigma }_+)^8$$
in $`[\nu _k,\tau _k]`$, and (15.27) follows. Since $`|x|1`$ and $`|\stackrel{~}{x}|1.1`$, cf. (15.26), we have
$$|\stackrel{~}{x}^2x^2|6e^{2k}(1+\mathrm{\Sigma }_+)^8,$$
so that
$$|\mathrm{\Sigma }_{}^2(1\mathrm{\Sigma }_+^2)x^2|12e^{2k}(1+\mathrm{\Sigma }_+)^9$$
in the interval $`[\nu _k,\tau _k]`$. $`\mathrm{}`$
Let us introduce
(15.31)
$$\eta (\tau )=2\xi (\tau )=2_{\tau _k}^\tau g(s)๐s+2\varphi _k,$$
where $`g=3(N_2+N_3)2(1+\mathrm{\Sigma }_+)\stackrel{~}{x}\stackrel{~}{y}=g_1+g_2`$. The reason we study $`\eta `$ instead of $`\xi `$ is that the trigonometric expression we will be interested in is $`\mathrm{sin}^2(\xi )`$, which has a period of length $`\pi `$, cf. Lemma 15.9. In the proof of Lemma 15.10, it is shown that, in the interval $`[\nu _k,\tau _k]`$, the first term appearing in $`g`$ is much greater than the second. We can thus consider functions of $`\tau `$ in the interval $`[\nu _k,\tau _k]`$ to be functions of $`\eta `$. We will mainly be interested in considering an interval $`[\eta _0,\eta _0+2\pi ]`$ at a time, so that we will only need to estimate the variation of the relevant expressions during one such period.
###### Lemma 15.10.
Let $`\eta _{1,k}=\eta (\sigma _k)`$ and $`\eta _{2,k}=\eta (\nu _k)`$. If $`[\eta _1,\eta _1+2\pi ][\eta _{1,k},\eta _{2,k}]`$ and $`\eta _a,\eta _b[\eta _1,\eta _1+2\pi ]`$, then for $`k`$ large enough
(15.32)
$$e^{6\pi /ฯต}\frac{(N_2+N_3)(\eta _a)}{(N_2+N_3),(\eta _b)}e^{6\pi /ฯต},$$
(15.33)
$$\frac{1}{2}\frac{1+\mathrm{\Sigma }_+(\eta _a)}{1+\mathrm{\Sigma }_+(\eta _b)}2$$
and
(15.34)
$$|g_1|/2|g|2|g_1|.$$
Proof. Because of Lemma 15.8,
(15.35)
$$\frac{1+\mathrm{\Sigma }_+}{N_2+N_3}\frac{1+\mathrm{\Sigma }_+}{2(N_2N_3)^{1/2}}=(N_2N_3)^{1/2}\frac{1+\mathrm{\Sigma }_+}{2N_2N_3}$$
$$\frac{1}{2ฯต}\left(\frac{N_2N_3}{(N_2N_3)(u_k)}\right)^{1/2}(N_2N_3)^{1/2}(u_k)\frac{1}{2ฯต^{1/2}}e^{2k}$$
in the interval $`[\nu _k,\tau _k]`$. By (15.26) we may assume $`\stackrel{~}{x}^2+\stackrel{~}{y}^22`$ in $`[\nu _k,\tau _k]`$. Combining this fact with (15.35) yields (15.34) in $`[\nu _k,\tau _k]`$. Thus, $`d\eta /d\tau <0`$ in that interval. We have
$$|\frac{d(N_2+N_3)}{d\eta }|=|\frac{1}{2g}((q+2\mathrm{\Sigma }_+)(N_2+N_3)+2\sqrt{3}\mathrm{\Sigma }_{}(N_2N_3))|$$
$$\frac{1}{2}(3\gamma 2)\mathrm{\Omega }+|\mathrm{\Sigma }_+^2+(1\mathrm{\Sigma }_+^2)\stackrel{~}{x}^2+\mathrm{\Sigma }_+|+2\frac{|\stackrel{~}{x}\stackrel{~}{y}|}{|g|}(1\mathrm{\Sigma }_+^2)6(1+\mathrm{\Sigma }_+)+8\frac{1+\mathrm{\Sigma }_+}{N_2+N_3},$$
so that
$$|\frac{1}{N_2+N_3}\frac{d(N_2+N_3)}{d\eta }|6\frac{1+\mathrm{\Sigma }_+}{N_2+N_3}+8\frac{1+\mathrm{\Sigma }_+}{(N_2+N_3)^2}6\frac{1+\mathrm{\Sigma }_+}{N_2+N_3}+2\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\frac{3}{ฯต}$$
in $`[\nu _k,\sigma _k]`$ for $`k`$ large, by Lemma 15.8 and (15.35). If $`N_2+N_3`$ has a maximum in $`\eta _{\mathrm{max}}[\eta _1,\eta _1+2\pi ]`$ and a minimum in $`\eta _{\mathrm{min}}`$, we get
$$\frac{(N_2+N_3)(\eta _{\mathrm{max}})}{(N_2+N_3)(\eta _{\mathrm{min}})}e^{6\pi /ฯต},$$
and (15.32) follows. We also need to know how much $`1+\mathrm{\Sigma }_+`$ varies over one period. By (2.4)
$$(1+\mathrm{\Sigma }_+)^{}=(2\mathrm{\Sigma }_+^2+2\mathrm{\Sigma }_{}^22)(1+\mathrm{\Sigma }_+)+f_1,$$
where $`f_1`$ is an expression that can be estimated as in (15.21), so that we in $`[\nu _k,\tau _k]`$ have
$$|\frac{(1+\mathrm{\Sigma }_+)^{}}{1+\mathrm{\Sigma }_+}|2(1\mathrm{\Sigma }_+^2)(1+\stackrel{~}{x}^2)+(1+\mathrm{\Sigma }_+)13(1+\mathrm{\Sigma }_+),$$
for $`k`$ large enough. Thus,
(15.36)
$$|\frac{1}{1+\mathrm{\Sigma }_+}\frac{d(1+\mathrm{\Sigma }_+)}{d\eta }|\frac{10(1+\mathrm{\Sigma }_+)}{N_2+N_3},$$
so that (15.33) holds if $`k`$ is big enough and $`|\eta _a\eta _b|2\pi `$ by (15.35). $`\mathrm{}`$
###### Lemma 15.11.
Let $`\sigma _k`$ and $`\tau _k`$ be as above. Then if $`k`$ is large enough,
$$\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\frac{1}{ฯต}e^{c_ฯตk}$$
in $`[\sigma _k,\tau _k]`$ where $`c_ฯต>0`$.
Proof. Observe that similarly to the proof of Lemma 15.7, we have
$$k_{\nu _k}^{\sigma _k}(1+\mathrm{\Sigma }_+)๐\tau =_{\eta _{1,k}}^{\eta _{2,k}}\frac{(1+\mathrm{\Sigma }_+)}{2g}๐\eta .$$
The contribution from one period in $`\eta `$ is negligible, by (15.35) and (15.34). Compare this integral with
$$_{\eta _{1,k}}^{\eta _{2,k}}\frac{\mathrm{\Sigma }_{}^2}{g}๐\eta =_{\eta _{1,k}}^{\eta _{2,k}}\frac{(1\mathrm{\Sigma }_+^2)x^2}{g}๐\eta +_{\eta _{1,k}}^{\eta _{2,k}}\frac{\mathrm{\Sigma }_{}^2(1\mathrm{\Sigma }_+^2)x^2}{g}๐\eta =I_{1,k}+I_{2,k}.$$
Now,
$$|I_{2,k}|e^k_{\eta _{1,k}}^{\eta _{2,k}}\frac{1+\mathrm{\Sigma }_+}{g}๐\eta $$
by (15.25). Consider an interval $`[\eta _1,\eta _1+2\pi ]`$. Estimate, letting $`\eta _a`$ and $`\eta _b`$ be the minimum and maximum of $`\mathrm{\Sigma }_+`$ respectively, and $`\eta _{\mathrm{min}}`$, $`\eta _{\mathrm{max}}`$ the min and max for $`g_1`$ in this interval,
$$_{\eta _1}^{\eta _1+2\pi }\frac{(1\mathrm{\Sigma }_+^2)x^2}{g}๐\eta _{\eta _1}^{\eta _1+2\pi }\frac{(1+\mathrm{\Sigma }_+)x^2}{g}๐\eta =_{\eta _1}^{\eta _1+2\pi }\frac{(1+\mathrm{\Sigma }_+)\mathrm{sin}^2(\eta /2)}{g}๐\eta $$
$$\pi \frac{1+\mathrm{\Sigma }_+(\eta _a)}{2|g_1(\eta _{\mathrm{max}})|}\frac{\pi }{2}e^{6\pi /ฯต}\frac{1+\mathrm{\Sigma }_+(\eta _a)}{|g_1(\eta _{\mathrm{min}})|}\frac{\pi }{4}e^{6\pi /ฯต}\frac{1+\mathrm{\Sigma }_+(\eta _b)}{|g_1(\eta _{\mathrm{min}})|}=$$
$$=\frac{1}{8}e^{6\pi /ฯต}_{\eta _1}^{\eta _1+2\pi }\frac{1+\mathrm{\Sigma }_+(\eta _b)}{|g_1(\eta _{\mathrm{min}})|}๐\eta \frac{1}{16}e^{6\pi /ฯต}_{\eta _1}^{\eta _1+2\pi }\frac{1+\mathrm{\Sigma }_+(\eta )}{g(\eta )}๐\eta ,$$
where we have used (15.32), (15.33) and (15.34). Assuming, without loss of generality, that $`\eta _{2,k}\eta _{1,k}`$ is an integer multiple of $`2\pi `$, we get
$$_{\nu _k}^{\sigma _k}2\mathrm{\Sigma }_{}^2๐\tau =_{\eta _{1,k}}^{\eta _{2,k}}\frac{\mathrm{\Sigma }_{}^2}{g}๐\eta =I_{1,k}+I_{2,k}\left(\frac{1}{16}e^{6\pi /ฯต}e^k\right)_{\eta _{1,k}}^{\eta _{2,k}}\frac{1+\mathrm{\Sigma }_+(\eta )}{g(\eta )}๐\eta $$
$$\frac{1}{20}e^{6\pi /ฯต}_{\eta _{1,k}}^{\eta _{2,k}}\frac{1+\mathrm{\Sigma }_+(\eta )}{g(\eta )}๐\eta =\frac{1}{10}e^{6\pi /ฯต}_{\nu _k}^{\sigma _k}(1+\mathrm{\Sigma }_+)๐\tau \frac{k}{10}e^{6\pi /ฯต}=c_ฯตk$$
for $`k`$ large enough and the lemma follows from (15.24). $`\mathrm{}`$
The following corollary summarizes the estimates that make the order of magnitude calculus well defined.
###### Corollary 15.2.
Let $`\sigma _k`$ and $`\tau _k`$ be as above. Then
(15.37)
$$\frac{1+\mathrm{\Sigma }_+}{(N_2+N_3)^2}\frac{1}{ฯต}e^{c_ฯตk},$$
(15.38)
$$\frac{1+\mathrm{\Sigma }_+}{N_2+N_3}e^{2k}$$
and
(15.39)
$$1e^{2k}\frac{g}{g_1}1+e^{2k}$$
in $`[\sigma _k,\tau _k]`$ for $`k`$ large enough.
Proof. Observe that by Lemma 15.11,
$$\frac{1+\mathrm{\Sigma }_+}{(N_2+N_3)^2}\frac{1+\mathrm{\Sigma }_+}{N_2N_3}\frac{1}{ฯต}e^{c_ฯตk}$$
and
$$\frac{1+\mathrm{\Sigma }_+}{N_2+N_3}\frac{1+\mathrm{\Sigma }_+}{2(N_2N_3)^{1/2}}ฯต^{1/2}e^{2k}\frac{1}{2ฯต}e^{c_ฯตk},e^{2k}$$
for $`k`$ large enough, cf. (15.35). We have
$$\frac{g}{g_1}=1+\frac{2(1+\mathrm{\Sigma }_+)\stackrel{~}{x}\stackrel{~}{y}}{3(N_2+N_3)}.$$
By (15.26) and the above estimates, we get (15.39) for $`k`$ large enough. $`\mathrm{}`$
The interval we will work with from now on is $`[\sigma _k,\tau _k]`$. Let $`\eta `$ be defined as in (15.31), but define $`\eta _{1,k}=\eta (\tau _k)`$ and $`\eta _{2,k}=\eta (\sigma _k)`$. We need to improve the estimates of the variation of $`1+\mathrm{\Sigma }_+`$ and $`N_2+N_3`$ during a period contained in $`[\eta _{1,k},\eta _{2,k}]`$.
###### Lemma 15.12.
Consider an interval $`=[\eta _1,\eta _1+2\pi ][\eta _{1,k},\eta _{2,k}]`$, where $`\eta _{1,k}=\eta (\tau _k)`$ and $`\eta _{2,k}=\eta (\sigma _k)`$. Let $`\eta _a`$ and $`\eta _b`$ correspond to the max and min of $`1+\mathrm{\Sigma }_+`$ in $``$, and let $`\eta _{\mathrm{max}}`$ and $`\eta _{\mathrm{min}}`$ correspond to the max and min of $`N_2+N_3`$ in the same interval. Then,
(15.40)
$$|\mathrm{\Sigma }_+(\eta _b)\mathrm{\Sigma }_+(\eta _a)|\frac{40\pi (1+\mathrm{\Sigma }_+(\eta _b))^2}{(N_2+N_3)(\eta _{\mathrm{max}})}$$
and
(15.41)
$$\frac{(N_2+N_3)(\eta _{\mathrm{max}})}{(N_2+N_3)(\eta _{\mathrm{min}})}\mathrm{exp}(\frac{20\pi }{ฯต}\mathrm{exp}(c_ฯตk)).$$
Proof. The derivation of (15.36) is still valid, so that
$$|\frac{1}{1+\mathrm{\Sigma }_+}\frac{d(1+\mathrm{\Sigma }_+)}{d\eta }|\frac{10(1+\mathrm{\Sigma }_+)}{N_2+N_3}.$$
By (15.38) we conclude that $`(1+\mathrm{\Sigma }_+(\eta _a))/(1+\mathrm{\Sigma }_+(\eta _b))`$ can be chosen to be arbitrarily close to one by choosing $`k`$ large enough. Now,
$$\frac{1}{N_2+N_3}\frac{d(N_2+N_3)}{d\eta }=\frac{1}{N_2+N_3}\frac{1}{2g}\frac{d(N_2+N_3)}{d\tau }=$$
$$=\frac{1}{N_2+N_3}\frac{1}{2g}((q+2\mathrm{\Sigma }_+)(N_2+N_3)+2\sqrt{3}\mathrm{\Sigma }_{}(N_2N_3))=\frac{q+2\mathrm{\Sigma }_+}{2g}+\frac{4(1\mathrm{\Sigma }_+^2)\stackrel{~}{x}\stackrel{~}{y}}{2(N_2+N_3)g},$$
and consequently
$$|\frac{1}{N_2+N_3}\frac{d(N_2+N_3)}{d\eta }|\frac{10}{ฯต}e^{c_ฯตk}.$$
Equation (15.41) follows, and the relative variation of $`N_2+N_3`$ during one period can be chosen arbitrarily small. Finally,
$$|\mathrm{\Sigma }_+(\eta _b)\mathrm{\Sigma }_+(\eta _a)|=(1+\mathrm{\Sigma }_+(\eta _b))|\frac{1+\mathrm{\Sigma }_+(\eta _a)}{1+\mathrm{\Sigma }_+(\eta _b)}1|$$
$$\frac{30\pi (1+\mathrm{\Sigma }_+(\eta _b))^2}{(N_2+N_3)(\eta _{\mathrm{min}})}$$
by (15.36) and the above observations. We may also change $`\eta _{\mathrm{min}}`$ to $`\eta _{\mathrm{max}}`$ at the cost of increasing the constant. $`\mathrm{}`$
As has been stated earlier, the goal of this section is to prove that the conditions of Lemma 15.2 are never met. We do this by deducing a contradiction from the consequences of that lemma. On the one hand, we have a rough picture of how the solution behaves in $`[\sigma _k,\tau _k]`$ by Lemma 15.9, Lemma 15.12 and Corollary 15.2. On the other hand, we know that, since $`r(\sigma _k)r(\tau _k)=k`$,
(15.42)
$$k=_{\sigma _k}^{\tau _k}(\frac{1}{4}(3\gamma 2)\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\mathrm{\Sigma }_+)๐\tau =\alpha _k+_{\eta _{1,k}}^{\eta _{2,k}}\frac{\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\mathrm{\Sigma }_+}{2g}๐\eta .$$
We will use our knowledge of the behaviour of the solution in $`[\sigma _k,\tau _k]`$ to prove that (15.42) is false. Observe that $`\eta _{1,k}<\eta _{2,k}`$, and that the contribution from one period is negligible, cf. Corollary 15.2. Also, $`\alpha _k0`$ as $`k\mathrm{}`$ so that we may ignore it. We will prove that for $`k`$ great enough, the integral of $`(\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\mathrm{\Sigma }_+)/(2g)`$ over a suitably chosen period is positive. From here on, we consider an interval $`[\eta _1,\eta _1+2\pi ]`$ which, excepting intervals of length less than a period at each end of $`[\eta _{1,k},\eta _{2,k}]`$, we can assume to be of the form $`[\pi /2,3\pi /2]`$. There is however one thing that should be kept in mind; when translating the $`\eta `$-variable by $`2m\pi `$ the $`\xi `$-variable is translated by $`m\pi `$. In other words, there is a sign involved, and in order to keep track of it we write out the details. By the above observations we have.
###### Lemma 15.13.
For each $`k`$ there are integers $`m_{1,k}`$ and $`m_{2,k}`$ such that
(15.43)
$$k=\beta _k+_{\pi /2+2m_{1,k}\pi }^{3\pi /2+2m_{2,k}\pi }\frac{\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\mathrm{\Sigma }_+}{2g}๐\eta ,$$
where $`\beta _k0`$ as $`k\mathrm{}`$, and
$$\eta _{1,k}\pi /2+2m_{1,k}\pi \eta _{1,k}+2\pi ,\eta _{2,k}2\pi 3\pi /2+2m_{2,k}\pi \eta _{2,k}.$$
Consider now an interval
$$[\pi /2+2m\pi ,3\pi /2+2m\pi ][\pi /2+2m_{1,k},3\pi /2+2m_{2,k}\pi ],$$
where $`m`$ is an integer, and make the substitution
$$\stackrel{~}{\eta }=\eta 2m\pi ,\stackrel{~}{\xi }=\xi m\pi $$
in that interval. Compute
$$\mathrm{\Sigma }_+^2+(1\mathrm{\Sigma }_+^2)x^2+\mathrm{\Sigma }_+=(1+\mathrm{\Sigma }_+)(\mathrm{\Sigma }_++(1\mathrm{\Sigma }_+)\frac{1}{2}(1\mathrm{cos}\eta ))=$$
$$=(1+\mathrm{\Sigma }_+)(\frac{1}{2}(1+\mathrm{\Sigma }_+)\frac{1}{2}(1\mathrm{\Sigma }_+)\mathrm{cos}\stackrel{~}{\eta })=\frac{1}{2}(1+\mathrm{\Sigma }_+)((1+\mathrm{\Sigma }_+)(1\mathrm{\Sigma }_+)\mathrm{cos}\stackrel{~}{\eta }).$$
This expression is the relevant part of the numerator of the integrand in the right hand side of (15.43). There is a drift term yielding a positive contribution to the integral, but the oscillatory term is arbitrarily much greater by Lemma 15.6. The interval $`[\pi /2,3\pi /2]`$ was not chosen at random. By considering the above expression, one concludes that the oscillatory term is negative in $`[\pi /2,\pi /2]`$ and positive in $`[\pi /2,3\pi /2]`$. As far as obtaining a contradiction goes, the first interval is thus bad and the second good. In order to estimate the integral over a period, the natural thing to do is then to make a substitution in the interval $`[\pi /2,3\pi /2]`$, so that it becomes an integral over the interval $`[\pi /2,\pi /2]`$. It is then important to know how the different expressions vary with $`\eta `$. We will prove a lemma saying that $`\mathrm{\Sigma }_+`$ roughly increases with $`\eta `$, and it will turn out to be useful that $`\mathrm{\Sigma }_+`$ is greater in the good part than in the bad. Let
(15.44)
$$J=_{\pi /2+2m\pi }^{3\pi /2+2m\pi }\frac{\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\mathrm{\Sigma }_+}{2g}๐\eta =\frac{1}{2}_{\pi /2}^{3\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+2m\pi ))^2}{2g(\stackrel{~}{\eta }+2m\pi )}๐\stackrel{~}{\eta }$$
$$\frac{1}{2}_{\pi /2}^{3\pi /2}\frac{(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+2m\pi ))\mathrm{cos}\stackrel{~}{\eta }}{2g(\stackrel{~}{\eta }+2m\pi )}๐\stackrel{~}{\eta }+$$
$$+_{\pi /2}^{3\pi /2}\frac{\mathrm{\Sigma }_{}^2(\stackrel{~}{\eta }+2m\pi )(1\mathrm{\Sigma }_+(\stackrel{~}{\eta }+2m\pi ))^2x^2(\stackrel{~}{\eta }+2m\pi )}{2g(\stackrel{~}{\eta }+2m\pi )}๐\stackrel{~}{\eta }=J_1+J_2+J_3.$$
If we can prove that $`J`$ is positive regardless of $`m`$ we are done, since $`J`$ positive contradicts (15.43). The integral $`J_1`$ is positive, and because the relative variation of the integrand can be chosen arbitrarily small by choosing $`k`$ large enough, $`J_1`$ is of the order of magnitude
(15.45)
$$\frac{(1+\mathrm{\Sigma }_+)^2}{N_2+N_3}.$$
If negative terms in $`J_2`$ and $`J_3`$ of the orders of magnitude
(15.46)
$$\frac{(1+\mathrm{\Sigma }_+)^3}{(N_2+N_3)^2}$$
or
(15.47)
$$\frac{(1+\mathrm{\Sigma }_+)^3}{(N_2+N_3)^3}$$
occur, we may ignore them by (15.38) and (15.37). By (15.25), $`J_3`$ may be ignored. Observe that the largest integrand is the one appearing in $`J_2`$. However, it oscillates. Considering (15.44), one can see that writing out arguments such as $`\stackrel{~}{\eta }+2m\pi `$ does not make things all that much clearer. For that reason, we introduce the following convention.
###### Convention 15.1.
By $`\mathrm{\Sigma }_+(\stackrel{~}{\eta })`$ and $`\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi )`$, we will mean $`\mathrm{\Sigma }_+(\stackrel{~}{\eta }+2m\pi )`$ and $`\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi +2m\pi )`$ respectively, and similarly for all expressions in the variables of Wainwright and Hsu. However, trigonometric expressions should be read as stated. Thus $`\mathrm{cos}(\stackrel{~}{\eta }/2)`$ means just that and not $`\mathrm{cos}(\stackrel{~}{\eta }/2+m\pi )`$.
###### Definition 15.1.
Consider an integral expression
$$I=_{\pi /2}^{3\pi /2}f(\stackrel{~}{\eta })๐\stackrel{~}{\eta }.$$
Then we say that $`I`$ is less than or equal to zero up to order of magnitude, if
$$I_{\pi /2}^{3\pi /2}g(\stackrel{~}{\eta })๐\stackrel{~}{\eta },$$
where $`g`$ satisfies a bound
$$gC_1\frac{(1+\mathrm{\Sigma }_+)^3}{(N_2+N_3)^2}+C_2\frac{(1+\mathrm{\Sigma }_+)^3}{(N_2+N_3)^3},$$
for $`k`$ large enough, where $`C_1`$ and $`C_2`$ are positive constants independent of $`k`$. We write $`I0`$. The definition of $`I0`$ is similar. We also define the concept similarly if the interval of integration is different.
We will use the same terminology more generally in inequalities between functions, if those inequalities, when inserted into the proper integrals, yield inequalities in the sense of the definition above. We will write $``$ if the error is of negligible order of magnitude.
###### Lemma 15.14.
If $`J_2`$ as defined above satisfies $`J_20`$, then $`J`$ is non-negative for $`k`$ large enough.
Proof. Under the assumptions of the lemma, we have
$$J\frac{1}{2}_{\pi /2}^{3\pi /2}\frac{(1+\mathrm{\Sigma }_+)^2}{2g}๐\stackrel{~}{\eta }_{\pi /2}^{3\pi /2}(C_1\frac{(1+\mathrm{\Sigma }_+)^3}{(N_2+N_3)^2}+C_2\frac{(1+\mathrm{\Sigma }_+)^3}{(N_2+N_3)^3})๐\stackrel{~}{\eta }+$$
$$+_{\pi /2}^{3\pi /2}\frac{\mathrm{\Sigma }_{}^2(1\mathrm{\Sigma }_+^2)x^2}{2g}๐\stackrel{~}{\eta }.$$
By Corollary 15.2, Lemma 15.12 and (15.25), we conclude that for $`k`$ large enough, $`J`$ is positive. $`\mathrm{}`$
The following lemma says that $`\mathrm{\Sigma }_+`$ almost increases with $`\stackrel{~}{\eta }`$.
###### Lemma 15.15.
Let $`\pi /2\stackrel{~}{\eta }_a\stackrel{~}{\eta }_b3\pi /2`$. Then
$$\mathrm{\Sigma }_+(\stackrel{~}{\eta }_b)\mathrm{\Sigma }_+(\stackrel{~}{\eta }_a)(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }_{\mathrm{min}}))^8,$$
where $`\stackrel{~}{\eta }_{\mathrm{min}}`$ corresponds to the minimum of $`1+\mathrm{\Sigma }_+`$ in $`[\pi /2,3\pi /2]`$.
Proof. We have
$$\mathrm{\Sigma }_+^{}\frac{3}{2}(2\gamma )\mathrm{\Omega },$$
so that
$$\frac{d\mathrm{\Sigma }_+}{d\stackrel{~}{\eta }}\frac{3}{2}(2\gamma )\frac{\mathrm{\Omega }}{2g}.$$
Using (15.21), (15.38) and Lemma 15.12, we conclude that
$$\frac{d\mathrm{\Sigma }_+}{d\stackrel{~}{\eta }}\frac{1}{2\pi }(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }_{\mathrm{min}}))^8.$$
The lemma follows. $`\mathrm{}`$
###### Lemma 15.16.
If
$$I=_{\pi /2}^{3\pi /2}\frac{1+\mathrm{\Sigma }_+}{g}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }$$
satisfies $`I0`$, then $`J_20`$.
Proof. Consider
$$J_2=_{\pi /2}^{3\pi /2}\frac{(1\mathrm{\Sigma }_+^2)\mathrm{cos}\stackrel{~}{\eta }}{4g}๐\stackrel{~}{\eta }=_{\pi /2}^{3\pi /2}\frac{(\mathrm{\Sigma }_+(3\pi /2)\mathrm{\Sigma }_+)(1+\mathrm{\Sigma }_+)}{4g}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }+$$
$$+(1\mathrm{\Sigma }_+(3\pi /2))_{\pi /2}^{3\pi /2}\frac{1+\mathrm{\Sigma }_+}{4g}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }.$$
The first integral is negligible by (15.40). The lemma follows. $`\mathrm{}`$
###### Lemma 15.17.
If
$$I_1=_{\pi /2}^{\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))(g_1(\stackrel{~}{\eta })g_1(\stackrel{~}{\eta }+\pi ))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }$$
satisfies $`I_10`$, then $`J_20`$.
Proof. We have
$$I=_{\pi /2}^{3\pi /2}\frac{1+\mathrm{\Sigma }_+}{g}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }=_{\pi /2}^{\pi /2}\frac{1+\mathrm{\Sigma }_+}{g}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }+_{\pi /2}^{3\pi /2}\frac{1+\mathrm{\Sigma }_+}{g}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }.$$
Make the substitution $`\chi =\stackrel{~}{\eta }+\pi `$ in the second integral;
$$_{\pi /2}^{\pi /2}\frac{1+\mathrm{\Sigma }_+(\chi +\pi )}{g(\chi +\pi )}\mathrm{cos}(\chi +\pi )(d\chi )=_{\pi /2}^{\pi /2}\frac{1+\mathrm{\Sigma }_+(\chi +\pi )}{g(\chi +\pi )}\mathrm{cos}(\chi )๐\chi .$$
Thus,
$$I=_{\pi /2}^{\pi /2}\left(\frac{1+\mathrm{\Sigma }_+(\stackrel{~}{\eta })}{g(\stackrel{~}{\eta })}\frac{1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi )}{g(\stackrel{~}{\eta }+\pi )}\right)\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }=$$
$$=_{\pi /2}^{\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi ))g(\stackrel{~}{\eta })(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))g(\stackrel{~}{\eta }+\pi )}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }.$$
But
$$(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi ))g(\stackrel{~}{\eta })(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))g(\stackrel{~}{\eta }),$$
by Lemma 15.15, so that
(15.48)
$$I_{\pi /2}^{\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))(g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi ))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }.$$
Now,
$$g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )=g_1(\stackrel{~}{\eta })g_1(\stackrel{~}{\eta }+\pi )+g_2(\stackrel{~}{\eta })g_2(\stackrel{~}{\eta }+\pi ),$$
but since $`2xy=\mathrm{sin}\stackrel{~}{\eta }`$ and the error committed in replacing $`\stackrel{~}{x}`$ with $`x`$ and $`\stackrel{~}{y}`$ with $`y`$ is negligible by (15.27), we have
$$g_2(\stackrel{~}{\eta })g_2(\stackrel{~}{\eta }+\pi )(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))\mathrm{sin}\stackrel{~}{\eta }+(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi ))\mathrm{sin}(\stackrel{~}{\eta }+\pi )=$$
$$=(\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi ))\mathrm{\Sigma }_+(\stackrel{~}{\eta }))\mathrm{sin}\stackrel{~}{\eta }.$$
The corresponding contribution to the integral may consequently be neglected; the error in the integral will be of type (15.47) by (15.40). Consequently, if
$$I_1=_{\pi /2}^{\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))(g_1(\stackrel{~}{\eta })g_1(\stackrel{~}{\eta }+\pi ))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }$$
satisfies $`I_10`$, then $`I0`$ by (15.48), so that the lemma follows by Lemma 15.16. $`\mathrm{}`$
Let
$$h_1(\stackrel{~}{\eta })=g_1(\stackrel{~}{\eta })g_1(\stackrel{~}{\eta }+\pi ).$$
We estimate $`h_1`$ by estimating the derivative. We have $`h_1(\pi /2)=0`$.
###### Lemma 15.18.
Let $`h_1`$ be as above. In the interval $`[\pi /2,\pi /2]`$, we have
(15.49)
$$\frac{dh_1}{d\stackrel{~}{\eta }}3\left(\frac{1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta })}{g(\stackrel{~}{\eta })}+\frac{1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi )}{g(\stackrel{~}{\eta }+\pi )}\right)\mathrm{sin}\stackrel{~}{\eta }.$$
Proof. Compute
$$\frac{dh_1}{d\stackrel{~}{\eta }}(\stackrel{~}{\eta })=\frac{dg_1}{d\stackrel{~}{\eta }}(\stackrel{~}{\eta })+\frac{dg_1}{d\stackrel{~}{\eta }}(\stackrel{~}{\eta }+\pi ).$$
But
$$\frac{dg_1}{d\stackrel{~}{\eta }}=\frac{3}{2g}((q+2\mathrm{\Sigma }_+)(N_2+N_3)+2\sqrt{3}\mathrm{\Sigma }_{}(N_2N_3))=$$
$$=\frac{1}{2}(q+2\mathrm{\Sigma }_+)\frac{gg_2}{g}3\sqrt{3}\frac{\mathrm{\Sigma }_{}(N_2N_3)}{g}.$$
Observe that $`x`$ and $`y`$ are trigonometric expressions, and that
$$2x(\stackrel{~}{\eta }+2m\pi )y(\stackrel{~}{\eta }+2m\pi )=2\mathrm{sin}(\stackrel{~}{\eta }/2+m\pi )\mathrm{cos}(\stackrel{~}{\eta }/2+m\pi )=\mathrm{sin}\stackrel{~}{\eta }.$$
We have
$$\sqrt{3}\mathrm{\Sigma }_{}(N_2N_3)2(1\mathrm{\Sigma }_+^2)xy=(1\mathrm{\Sigma }_+^2)\mathrm{sin}\stackrel{~}{\eta },$$
so that
$$\frac{dg_1}{d\stackrel{~}{\eta }}(\frac{1}{4}(3\gamma 2)\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\mathrm{\Sigma }_+)\frac{g_2}{g}(\frac{1}{4}(3\gamma 2)\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2+\mathrm{\Sigma }_+)$$
$$\frac{3(1\mathrm{\Sigma }_+^2)\mathrm{sin}\stackrel{~}{\eta }}{g}.$$
The middle term and all terms involving $`\mathrm{\Omega }`$ may be ignored. Estimate
$$\mathrm{\Sigma }_+^2(\stackrel{~}{\eta })+\mathrm{\Sigma }_{}^2(\stackrel{~}{\eta })+\mathrm{\Sigma }_+(\stackrel{~}{\eta })+\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi )+\mathrm{\Sigma }_{}^2(\stackrel{~}{\eta }+\pi )+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi )$$
$$\mathrm{\Sigma }_+^2(\stackrel{~}{\eta })+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }))(\mathrm{sin}^2(\stackrel{~}{\eta }/2+m\pi )1/2)+\frac{1}{2}(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }))+\mathrm{\Sigma }_+(\stackrel{~}{\eta })+$$
$$+\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi )+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi ))(\mathrm{cos}^2(\stackrel{~}{\eta }/2+m\pi )1/2)+\frac{1}{2}(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi ))+$$
$$+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi )=\frac{1}{2}(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))^2+\frac{1}{2}(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi ))^2+$$
$$+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }))(\mathrm{sin}^2(\stackrel{~}{\eta }/2)1/2)+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi ))(\mathrm{cos}^2(\stackrel{~}{\eta }/2)1/2).$$
The first equality is a consequence of (15.25). Due to the fact that $`\stackrel{~}{\eta }[\pi /2,\pi /2]`$, we have $`\mathrm{cos}^2(\stackrel{~}{\eta }/2)1/20`$. Since $`\stackrel{~}{\eta }+\pi \stackrel{~}{\eta }`$ and $`\mathrm{\Sigma }_+`$ increases with $`\stackrel{~}{\eta }`$ up to order of magnitude according to Lemma 15.15, we have
$$1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi )1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }).$$
Consequently,
$$\frac{1}{2}(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))^2+\frac{1}{2}(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi ))^2+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }))(\mathrm{sin}^2(\stackrel{~}{\eta }/2)1/2)+$$
$$+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }+\pi ))(\mathrm{cos}^2(\stackrel{~}{\eta }/2)1/2)\frac{1}{2}(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))^2+\frac{1}{2}(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }+\pi ))^2+$$
$$+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }))(\mathrm{sin}^2(\stackrel{~}{\eta }/2)1/2)+(1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }))(\mathrm{cos}^2(\stackrel{~}{\eta }/2)1/2)0.$$
In other words, we have (15.49). Here the importance of the fact that $`\mathrm{\Sigma }_+`$ is greater in the good part than in the bad becomes apparent. $`\mathrm{}`$
###### Lemma 15.19.
Let $`I_1`$ be defined as above. Then $`I_10`$.
Proof. Let $`\stackrel{~}{\eta }_{\mathrm{max}}`$ and $`\stackrel{~}{\eta }_{\mathrm{min}}`$ correspond to the max and min of $`g`$ in the interval $`[\pi /2,3\pi /2]`$, and let $`\stackrel{~}{\eta }_a`$ and $`\stackrel{~}{\eta }_b`$ correspond to the max and min of $`\mathrm{\Sigma }_+`$, in the same interval. Observe that for $`\stackrel{~}{\eta }[\pi /2,3\pi /2]`$, we have
$$1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }_a)1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta })1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }_b).$$
In order not to obtain too complicated expressions, let us introduce the following terminology:
$$a_1=6\frac{1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }_b)}{g(\stackrel{~}{\eta }_{\mathrm{max}})}6\frac{1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta })}{g(\stackrel{~}{\eta })}6\frac{1\mathrm{\Sigma }_+^2(\stackrel{~}{\eta }_a)}{g(\stackrel{~}{\eta }_{\mathrm{min}})}=a_2\mathrm{and}$$
$$b_1=\frac{1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }_b)}{g^2(\stackrel{~}{\eta }_{\mathrm{max}})}\frac{1+\mathrm{\Sigma }_+(\stackrel{~}{\eta })}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\frac{1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }_a)}{g^2(\stackrel{~}{\eta }_{\mathrm{min}})}=b_2,$$
where $`\stackrel{~}{\eta }[\pi /2,3\pi /2]`$. Observe that
(15.50)
$$\underset{k\mathrm{}}{lim}\frac{a_1}{a_2}=\underset{k\mathrm{}}{lim}\frac{b_1}{b_2}=1,$$
by Corollary 15.2 and Lemma 15.12. Consider the interval $`[0,\pi /2]`$. By (15.49), we have
(15.51)
$$\frac{dh_1}{d\stackrel{~}{\eta }}a_1\mathrm{sin}\stackrel{~}{\eta },$$
so that
$$h_1(\stackrel{~}{\eta })=h_1(\pi /2)_{\stackrel{~}{\eta }}^{\pi /2}\frac{dh_1}{d\stackrel{~}{\eta }}๐\stackrel{~}{\eta }a_1\mathrm{cos}\stackrel{~}{\eta }$$
in the interval $`[0,\pi /2]`$. Now consider the interval $`[\pi /2,0]`$. We have
$$\frac{dh_1}{d\stackrel{~}{\eta }}a_2\mathrm{sin}\stackrel{~}{\eta }.$$
Consequently,
$$h_1(\stackrel{~}{\eta })=h_1(0)_{\stackrel{~}{\eta }}^0\frac{dh_1}{d\stackrel{~}{\eta }}๐\stackrel{~}{\eta }a_1+a_2(1\mathrm{cos}\stackrel{~}{\eta })$$
in the interval $`[\pi /2,0]`$. Estimate
$$_0^{\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))(g_1(\stackrel{~}{\eta })g_1(\stackrel{~}{\eta }+\pi ))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }=$$
$$=_0^{\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))h_1(\stackrel{~}{\eta })}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }_0^{\pi /2}\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}(a_1\mathrm{cos}^2\stackrel{~}{\eta })๐\stackrel{~}{\eta }$$
$$a_1b_1_0^{\pi /2}\mathrm{cos}^2\stackrel{~}{\eta }d\stackrel{~}{\eta }=\frac{\pi a_1b_1}{4}.$$
We also estimate
$$_{\pi /2}^0\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))(g_1(\stackrel{~}{\eta })g_1(\stackrel{~}{\eta }+\pi ))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }=$$
$$=_{\pi /2}^0\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))h_1(\stackrel{~}{\eta })}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }a_1_{\pi /2}^0\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }+$$
$$+a_2_{\pi /2}^0\frac{(1+\mathrm{\Sigma }_+(\stackrel{~}{\eta }))}{g(\stackrel{~}{\eta })g(\stackrel{~}{\eta }+\pi )}(1\mathrm{cos}\stackrel{~}{\eta })\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }a_1b_1_{\pi /2}^0\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }+$$
$$+a_2b_2_{\pi /2}^0(1\mathrm{cos}\stackrel{~}{\eta })\mathrm{cos}\stackrel{~}{\eta }d\stackrel{~}{\eta }a_1b_1+(1\frac{\pi }{4})a_2b_2.$$
Adding up, we conclude that
$$I_1(1+\pi /4)a_1b_1+(1\pi /4)a_2b_2=[(1+\pi /4)\frac{a_1b_1}{a_2b_2}+(1\pi /4)]a_2b_2,$$
which is negative for $`k`$ large enough by (15.50). Thus $`I_10`$. $`\mathrm{}`$
###### Theorem 15.1.
The conditions of Lemma 15.2 are never met.
Proof. If the conditions are met, then Lemma 15.13 follows, and also that it is false, by Lemmas 15.19, 15.17, 15.14 and (15.44). $`\mathrm{}`$
###### Corollary 15.3.
Let $`2/3<\gamma <2`$. For every $`ฯต>0`$ there is a $`\delta >0`$ such that if $`x`$ constitutes Bianchi IX initial data for (2.1)-(2.3) and
$$\underset{y๐}{inf}xy\delta $$
then
$$\underset{y๐}{inf}\mathrm{\Phi }(\tau ,x)yฯต$$
for all $`\tau 0`$, where $`\mathrm{\Phi }`$ is the flow of (2.1)-(2.3).
Proof. Assuming the contrary, there is an $`ฯต>0`$ and a sequence $`x_l๐`$ such that
$$\underset{y๐}{inf}\mathrm{\Phi }(s_l,x_l)yฯต$$
for some $`s_l0`$. Let $`\tau _l=0`$. Since $`d(\tau _l,x_l)0`$ and we can assume $`ฯต`$ is small enough that Proposition 14.1 is applicable, there must be an $`\eta >0`$ such that $`h(s_l,x_l)>\eta `$ for $`l`$ large enough, contradicting Theorem 15.1. $`\mathrm{}`$
###### Corollary 15.4.
Consider a generic Bianchi IX solution with $`2/3<\gamma <2`$. Then
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Omega }+N_1N_2+N_2N_3+N_1N_3)=0.$$
Proof. If $`h`$ does not converge to zero, then the conditions of Lemma 15.2 are met, since there for a generic solution is an $`\alpha `$-limit point on the Kasner circle by Proposition 13.1. Corollary 14.1 then yields the desired conclusion. $`\mathrm{}`$
Let $`๐`$ be the set of vacuum type I and II points as in Definition 1.6. By Corollary 15.4, a generic type IX solution with $`2/3<\gamma <2`$ converges to $`๐`$.
###### Corollary 15.5.
Let $`2/3<\gamma <2`$. The closure of $`_{\mathrm{IX}}`$ and the closure of $`๐ซ_{\mathrm{IX}}`$ do not intersect $`๐`$. Furthermore, the set of generic Bianchi IX points is open in the set of Bianchi IX points.
Remark. The closure of the Taub type IX points does intersect $`๐`$.
Proof. Assume there is a sequence $`x_l_{\mathrm{IX}}`$ such that $`x_lx๐`$. Let $`\tau _l=0`$. Observe that then $`d(x_l,\tau _l)0`$. By Theorem 15.1, there is for each $`ฯต>0`$ and for each $`L`$ an $`lL`$ such that $`h(\tau ,x_l)ฯต`$ for $`\tau \tau _l=0`$. By choosing $`L`$ large enough, we can assume $`\mathrm{\Omega }(\tau _l,x_l)`$ to be arbitrarily small and by choosing $`ฯต`$ small enough, we can assume that Proposition 14.1 is applicable. Consequently, we can assume $`\mathrm{\Omega }(\tau ,x_l)`$ to be as small as we wish for $`\tau (\mathrm{},\tau _l]`$, contradicting the fact that $`\mathrm{\Omega }(\tau ,x_l)1`$ as $`\tau \mathrm{}`$. The argument for $`๐ซ_{\mathrm{IX}}`$ is similar, since the $`\mathrm{\Omega }`$-coordinate of $`P_i^+(II)`$ is positive.
Consider now a generic point $`x`$ in the set of Bianchi IX points. There is a neighbourhood of $`x`$ that does not intersect the Taub points. Let us prove the similar statement for $`_{\mathrm{IX}}`$ and $`๐ซ_{\mathrm{IX}}`$. Assume there is a sequence $`x_l_{\mathrm{IX}}`$ such that $`x_lx`$. For each $`ฯต>0`$ there is a $`T0`$ such that $`d(T,\mathrm{\Phi }(T,x))ฯต/2`$, by Corollary 15.4. By continuity of the flow and the function $`d`$, we conclude that for $`l`$ large enough we have $`d(T,\mathrm{\Phi }(T,x_l))ฯต`$. Since $`\mathrm{\Phi }(T,x_l)_{\mathrm{IX}}`$, we get a contradiction to the first part of the lemma. Thus, there is an open neighbourhood of $`x`$ that does not intersect $`_{\mathrm{IX}}`$. The argument for $`๐ซ_{\mathrm{IX}}`$ is similar. $`\mathrm{}`$
###### Corollary 15.6.
Let $`2/3<\gamma <2`$. The closure of $`_{\mathrm{VII}_0}`$ and the closure of $`๐ซ_{\mathrm{VII}_0}`$ do not intersect $`๐`$. Furthermore, the generic Bianchi $`\mathrm{VII}_0`$ points are open in the set of Bianchi $`\mathrm{VII}_0`$ points.
Proof. The argument proving the first part is as in the Bianchi IX case, once one has checked that analogues of Proposition 14.1 and Theorem 15.1 hold in the Bianchi VI$`\mathrm{I}_0`$ case. The second part then follows as in the Bianchi IX case, using Proposition 10.2. $`\mathrm{}`$
## 16. Regularity of the set of non-generic points
Observe that the constraint (2.3) together with the additional assumption $`\mathrm{\Omega }0`$ defines a 5-dimensional submanifold of $`^6`$ which has a 4-dimensional boundary given by the vacuum points. We have the following.
###### Theorem 16.1.
Let $`2/3<\gamma <2`$. The sets $`_{\mathrm{II}},_{\mathrm{VII}_0}`$, $`_{\mathrm{IX}}`$, $`๐ซ_{\mathrm{VII}_0}`$ and $`๐ซ_{\mathrm{IX}}`$ are $`C^1`$ submanifolds of $`^6`$ of dimensions $`1`$, $`2`$, $`3`$, $`1`$ and $`2`$ respectively.
We prove this theorem at the end of this section. The idea is as follows. The only obstruction to e. g. $`_{\mathrm{II}}`$ being a $`C^1`$ submanifold, is if there is an open set $`O`$ containing $`F`$ and a sequence $`x_k_{\mathrm{II}}`$ such that $`x_kF`$, but each $`x_k`$ has to leave $`O`$ before it can converge to $`F`$. If there is such a sequence, we produce a sequence $`y_k_{\mathrm{II}}`$ such that the distance from $`y_k`$ to $`๐`$ converges to zero, contradicting Lemma 9.1. The argument is similar in the other cases.
We will need some results from . The theorem stated below is a special case of Theorem 6.2, p. 243.
###### Theorem 16.2.
In the differential equation
(16.1)
$$\xi ^{}=E\xi +G(\xi )$$
let $`G`$ be of class $`C^1`$ and $`G(0)=0,_\xi G(0)=0`$. Let $`E`$ have $`e>0`$ eigenvalues with positive real parts, $`d>0`$ eigenvalues with negative real parts and no eigenvalues with zero real part. Let $`\xi _t=\xi (t,\xi _0)`$ be the solution of (16.1) satisfying $`\xi (0,\xi _0)=\xi _0`$ and $`T^t`$ the corresponding map $`T^t(\xi _0)=\xi (t,\xi _0)`$. Then there exists a map $`R`$ of a neighbourhood of $`\xi =0`$ in $`\xi `$-space onto a neighbourhood of the origin in Euclidean $`(u,v)`$-space, where $`\mathrm{dim}(u)=d`$ and $`\mathrm{dim}(v)=e`$, such that $`R`$ is $`C^1`$ with non-vanishing Jacobian and $`RT^tR^1`$ has the form
(16.2)
$$\left(\begin{array}{c}u_t\\ v_t\end{array}\right)=\left(\begin{array}{c}e^{tP}u_0+U(t,u_0,v_0)\\ e^{tQ}v_0+V(t,u_0,v_0)\end{array}\right).$$
$`U,V`$ and their partial derivatives with respect to $`u_0,v_0`$ vanish at $`(u_0,v_0)=0`$. Furthermore $`V=0`$ if $`v_0=0`$ and $`U=0`$ if $`u_0=0`$. Finally $`e^P<1`$ and $`e^Q<1`$.
Let us begin by considering the local behaviour close to the fixed points.
###### Lemma 16.1.
Consider the critical point $`F`$. There is an open neighbourhood $`O`$ of $`F`$ in $`^6`$, and a 1-dimensional $`C^1`$ submanifold $`M_{\mathrm{II}}_{\mathrm{II}}`$ of $`O_{\mathrm{II}}`$, such that for each $`xO_{\mathrm{II}}`$, either $`xM_{\mathrm{II}}`$, or $`x`$ will leave $`O`$ as the flow of (2.1)-(2.3) is applied to $`x`$ in the negative time direction. Similarly, we get a 2-dimensional $`C^1`$ submanifold $`M_{\mathrm{VII}_0}`$ of $`O_{\mathrm{VII}_0}`$, and a 3-dimensional $`C^1`$ submanifold $`M_{\mathrm{IX}}`$ of $`O_{\mathrm{IX}}`$ with the same properties. Consider the critical point $`P_1^+(II)`$. We then have a similar situation. Give the neighbourhood corresponding to $`O`$ the name $`P`$, and use the letter $`N`$ instead of the letter $`M`$ to denote the relevant submanifolds. Then $`N_{\mathrm{VII}_0}`$ has dimension 1 and $`N_{\mathrm{IX}}`$ has dimension 2.
Proof. Observe that when $`\mathrm{\Omega }>0`$, we can consider (2.1)-(2.3) to be an unconstrained system of equations in five variables. Using the constraint (2.3) to express $`\mathrm{\Omega }`$ in terms of the other variables, we can ignore $`\mathrm{\Omega }`$ and consider the first five equations of (2.1) as a set of equations on an open submanifold of $`^5`$, defined by the condition $`\mathrm{\Omega }>0`$ (considering $`\mathrm{\Omega }`$ as a function of the other variables). In the Bianchi VI$`\mathrm{I}_0`$ case, we can consider the system to be unconstrained in four variables.
Let us first deal with the Bianchi VI$`\mathrm{I}_0`$ case. Consider the fixed point $`P_1^+(II)`$. Considering the Bianchi $`\mathrm{VII}_0`$ points with $`N_1,N_2>0`$ and $`N_3=0`$, the linearization has one eigenvalue with positive real part and three with negative real part, cf. . By a suitable translation of the variables, reversal of time, and a suitable definition of $`G`$ and $`E`$ in (16.1), we can consider a solution to (2.1)-(2.3) converging to $`P_1^+(II)`$ as $`\tau \mathrm{}`$ as a solution $`\xi `$ to (16.1) converging to $`0`$ as $`t\mathrm{}`$. $`E`$ has one eigenvalue with negative real part and three with positive real part, so that Theorem 16.2 yields a $`C^1`$ map $`R`$ of a neighbourhood of $`0`$ with non-vanishing Jacobian to a neighbourhood of the origin in $`^4`$, such that the flow takes the form (16.2) where $`u`$ and $`v^3`$.
Observe that since $`\xi =0`$ is a fixed point, there is a neighbourhood of that point such that the flow is defined for $`|t|1`$. There is also an open bounded ball $`B`$ centered at the origin in $`(u_0,v_0)`$-space such that $`U`$ and $`V`$ are defined in a neighbourhood $`N`$ of $`[1,1]\times B`$. Let $`a=e^P`$ and $`1/c=e^Q`$. For any $`ฯต>0`$, we can choose $`B`$ and then $`N`$ small enough that the norms of $`U,V`$ and their partial derivatives with respect to $`u`$ and $`v`$ are smaller than $`ฯต`$ in $`N`$. Assume $`B`$ and $`N`$ are such for some $`ฯต`$ satisfying
(16.3)
$$ฯต<\mathrm{min}\{\frac{c1}{2},\frac{1a}{2}\}.$$
Consider a solution $`\xi `$ to (16.1) such that $`R\xi (t)B`$ for all $`tT`$. Let $`(u_t,v_t)=R(\xi (t))`$ for $`tT`$. We wish to prove that $`v_t=0`$, and assume therefore that $`v_{t_0}0`$ for some $`t_0T`$. We have
$$v_{t_0+n}e^Qv_{t_0+n1}+V(1,v_{t_0+n1},u_{t_0+n1})$$
$$cv_{t_0+n1}ฯตv_{t_0+n1}\frac{1+c}{2}v_{t_0+n1},$$
where we have used (16.3), the fact that $`V`$ is zero when $`v_0=0`$, and the fact that $`(u_t,v_t)`$ remain in $`B`$ for $`tT`$. Thus,
$$v_{t_0+n}\left(\frac{1+c}{2}\right)^nv_{t_0},$$
which is irreconcilable with the fact that $`v_t`$ remains bounded.
If $`(u_{t_0},v_{t_0})B`$ and $`v_{t_0}=0`$, (16.2) yields $`v_{t_0+1}=0`$ and
$$u_{t_0+1}(a+\frac{1a}{2})u_{t_0}=\frac{1+a}{2}u_{t_0}.$$
Consequently, all points $`(u,v)B`$ with $`v=0`$ converge to $`(0,0)`$ as one applies the flow.
We are now in a position to go backwards in order to obtain the conclusions of the lemma. The set $`R^1(B)`$ will, after suitable operations, including non-unique extensions, turn into the set $`P`$ and $`R^1(\{v=0\}B)`$ turns into $`N_{\mathrm{VII}_0}`$. One can carry out a similar construction in the Bianchi IX case. Observe that one might then get a different $`P`$, but by taking the intersection we can assume them to be the same. The dimension of $`N_{\mathrm{IX}}`$ follows from a computation of the eigenvalues.
The argument concerning the fixed point $`F`$ is similar. $`\mathrm{}`$
Proof of Theorem 16.1. Let $`O`$, $`M_{\mathrm{II}}`$ and so on be as in the statement of Lemma 16.1. Observe that if there is a neighbourhood $`\stackrel{~}{O}O`$ of $`F`$ such that $`_{\mathrm{II}}\stackrel{~}{O}=M_{\mathrm{II}}\stackrel{~}{O}`$, then $`_{\mathrm{II}}`$ is a $`C^1`$ submanifold. The reason is that given any $`x_{\mathrm{II}}`$, there is a $`T`$ such that $`\mathrm{\Phi }(\tau ,x)\stackrel{~}{O}`$ for all $`\tau T`$. By Lemma 16.1, we conclude that $`\mathrm{\Phi }(T,x)M_{\mathrm{II}}`$. Then there is a neighbourhood $`O^{}\stackrel{~}{O}`$ of $`\mathrm{\Phi }(T,x)`$ such that $`O^{}_{\mathrm{II}}=O^{}M_{\mathrm{II}}`$. We thus get, for $`O^{}`$ suitably chosen, a $`C^1`$ map $`\psi :O^{}^6`$ with $`C^1`$ inverse, sending $`_{\mathrm{II}}O^{}`$ to a one dimensional hyperplane. If $`O^{}`$ is small enough, we can apply $`\mathrm{\Phi }(T,)`$ to it obtaining a neighbourhood of $`x`$. By the invariance of $`_{\mathrm{II}}`$, we have
$$\mathrm{\Phi }(T,O^{})_{\mathrm{II}}=\mathrm{\Phi }(T,O^{}_{\mathrm{II}}).$$
In other words, $`\mathrm{\Phi }(T,\psi ())`$ defines coordinates on $`\mathrm{\Phi }(T,O^{})`$ straightening out $`_{\mathrm{II}}`$. The arguments for the other cases are similar.
Let us now assume, in order to reach a contradiction, that there is a sequence $`x_k_{\mathrm{II}}O`$ such that $`x_kF`$ but $`x_kM_{\mathrm{II}}`$ for all $`k`$. If we let $`O^{}O`$ be a small enough ball containing $`F`$, we can assume that $`|N_i|^{}0`$ for $`i=1,2,3`$ in $`O^{}`$, cf. the proof of Lemma 4.2. For $`k`$ large enough, $`x_kO^{}`$ and applying the flow to them we obtain points $`y_k_{\mathrm{II}}O^{}`$. By choosing a suitable subsequence, we can assume that $`y_k`$ converges to a type I point $`y`$ which is not $`F`$. Given $`ฯต>0`$, there is a $`T`$ such that $`\mathrm{\Phi }(T,y)`$ is at distance less than $`ฯต/2`$ from $`๐`$. For $`k`$ large enough, $`\mathrm{\Phi }(T,y_k)_{\mathrm{II}}`$ will then be at distance less than $`ฯต`$ from $`๐`$. We get a contradiction to Lemma 9.1. The arguments for $`_{\mathrm{VII}_0}`$ and $`_{\mathrm{IX}}`$ are similar, due to Corollaries 15.6 and 15.5.
For $`๐ซ_{\mathrm{VII}_0}`$ and $`๐ซ_{\mathrm{IX}}`$, we need to modify the argument. Assume there is a sequence $`x_k๐ซ_{\mathrm{VII}_0}P`$ such that $`x_kP_1^+(II)`$, but $`x_kN_{\mathrm{VII}_0}`$ for all $`k`$. By choosing $`P^{}P`$ as a small enough ball, we can assume that $`|N_i|^{}0`$ in $`P^{}`$ for $`i=2,3`$, cf. the proof of Lemma 4.1. For $`k`$ large enough, $`x_kP^{}`$, and applying the flow to them we obtain points $`y_k๐ซ_{\mathrm{VII}_0}P^{}`$. By choosing a suitable subsequence, we can assume that $`y_k`$ converges to a type II point $`y`$ which is not $`P_1^+(II)`$. If $`y_{\mathrm{II}}`$, we can apply the same kind of reasoning as before, using Proposition 9.1 to get a contradiction to the consequences of Corollary 15.6. If $`y_{\mathrm{II}}`$ we get, by applying the flow to the points $`y_k`$, a sequence $`z_k๐ซ_{\mathrm{VII}_0}`$ converging to $`F`$. Applying the flow again, as before, we get a contradiction. The Bianchi IX case is similar using Corollary 15.5. $`\mathrm{}`$
## 17. Uniform convergence to the attractor
If $`x`$ constitutes initial data to (2.1)-(2.3) at $`\tau =0`$, then we denote the corresponding solution $`\mathrm{\Sigma }_+(\tau ,x)`$ and so on.
###### Proposition 17.1.
Let $`2/3<\gamma 2`$ and let $`K`$ be a compact set of Bianchi IX initial data. Then $`N_1N_2N_3`$ converges uniformly to zero on $`K`$. That is, for all $`ฯต>0`$ there is a $`T`$ such that
$$(N_1N_2N_3)(\tau ,x)ฯต$$
for all $`\tau T`$ and all $`xK`$.
Proof. Assume that $`N_1N_2N_3`$ does not converge to zero uniformly. Then there is an $`ฯต>0`$, a sequence $`\tau _k\mathrm{}`$ and $`x_kK`$ such that
$$(N_1N_2N_3)(\tau _k,x_k)ฯต.$$
We may assume, by choosing a convergent subsequence, that $`x_kx_{}`$ as $`k\mathrm{}`$. Because of the monotonicity of $`(N_1N_2N_3)(,x_k)`$, we conclude that
$$(N_1N_2N_3)(\tau ,x_k)ฯต.$$
for $`\tau [\tau _k,0]`$. Thus
$$(N_1N_2N_3)(\tau ,x_{})=\underset{k\mathrm{}}{lim}(N_1N_2N_3)(\tau ,x_k)ฯต$$
for all $`\tau 0`$. We have a contradiction. $`\mathrm{}`$
###### Corollary 17.1.
Let $`2/3<\gamma 2`$ and let $`K`$ be a compact set of Bianchi IX initial data. Then for every $`ฯต>0`$, there is a $`T`$ such that
$$\mathrm{\Omega }+\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^21+ฯต$$
for all $`xK`$ and $`\tau T`$.
Proof. As before. $`\mathrm{}`$
Consider
$$d=\mathrm{\Omega }+N_1N_2+N_2N_3+N_3N_1.$$
###### Proposition 17.2.
Let $`K`$ be a compact set of generic Bianchi IX initial data with $`2/3<\gamma <2`$. Then $`d`$ converges uniformly to zero on $`K`$.
Proof. Assume that $`d`$ does not converge to zero uniformly. Then there is an $`\eta >0`$, a sequence $`\tau _k\mathrm{}`$ and a sequence $`x_kK`$ such that
(17.1)
$$d(\tau _k,x_k)\eta .$$
We now prove that there is no sequence $`s_{k_n}`$ such that $`\tau _{k_n}s_{k_n}0`$ and
$$d(s_{k_n},x_{k_n})0.$$
Assume there is. By Theorem 15.1, there is no $`\delta >0`$ such that maximum of $`h(,x_{k_n})`$ in $`[\tau _{k_n},s_{k_n}]`$ exceeds $`\delta `$ for all $`n`$. For $`\delta `$ small enough, we can apply Proposition 14.1 to the interval $`[\tau _{k_n},s_{k_n}]`$ to conclude that for some $`n`$, $`\mathrm{\Omega }`$ cannot grow in very much in that interval either. We obtain a contradiction to (17.1) for $`\delta `$ small enough and $`n`$ big enough.
Thus there is an $`ฯต>0`$ such that
$$d(\tau ,x_k)ฯต$$
for all $`\tau [\tau _k,0]`$ and all $`k`$. Assume $`x_kx_{}`$. Then
$$d(\tau ,x_{})=\underset{k\mathrm{}}{lim}d(\tau ,x_k)ฯต>0$$
for all $`\tau 0`$. But $`x_{}`$ constitutes generic initial data. $`\mathrm{}`$
## 18. Existence of non-special $`\alpha `$-limit points on the Kasner circle
We know that there is an $`\alpha `$-limit point on the Kasner circle, but in order to prove curvature blow up we wish to prove the existence of a non-special $`\alpha `$-limit point on the Kasner circle.
###### Lemma 18.1.
Consider a generic Bianchi IX solution with $`2/3<\gamma <2`$. If it has a special point on the Kasner circle as an $`\alpha `$-limit point then it has an infinite number of $`\alpha `$-limit points on the Kasner circle.
Proof. By applying the symmetries, we can assume that there is an $`\alpha `$-limit point on the Kasner circle with $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(1,0)`$. Since the solution is not of Taub type, $`(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})`$ cannot converge to $`(1,0)`$ by Proposition 3.1. Thus there is an $`1>ฯต>0`$ such that for each $`T`$ there is a $`\tau T`$ such that $`1+\mathrm{\Sigma }_+(\tau )ฯต`$. Let $`\tau _k\mathrm{}`$ be such that $`\mathrm{\Sigma }_+(\tau _k)1`$.
Let $`\eta >0`$ satisfy $`\eta <ฯต`$. We wish to prove that there is a non-special $`\alpha `$-limit point on the Kasner circle with $`1+\mathrm{\Sigma }_+\eta `$. There is a sequence $`t_k\tau _k`$ such that $`1+\mathrm{\Sigma }_+(t_k)=\eta `$ and $`\mathrm{\Sigma }_+^{}(t_k)0`$ assuming $`k`$ is large enough. The condition on the derivative is possible to impose due to the fact that $`1+\mathrm{\Sigma }_+`$ eventually has to become greater than $`ฯต`$. Choosing a suitable subsequence of $`\{t_k\}`$, we get an $`\alpha `$-limit point which has to be a vacuum type I or II point by Corollary 15.4. If it is of type I, we get an $`\alpha `$-limit point on the Kasner circle with $`1+\mathrm{\Sigma }_+=\eta `$ and we are done. The $`\alpha `$-limit point cannot have $`N_1>0`$, because of the condition on the derivative, cf. the proof of Proposition 5.1. If it is of type II with $`N_2`$ or $`N_3`$ greater than zero, we can apply the flow to get a type II solution, call it $`x`$, of $`\alpha `$-limit points to the original solution. Since a type II solution with $`N_2`$ or $`N_3`$ greater than zero satisfies $`\mathrm{\Sigma }_+^{}<0`$, the $`\omega `$-limit point $`y`$ of $`x`$ must have $`1+\mathrm{\Sigma }_+<\eta `$. By Proposition 5.1 $`y๐ฆ_2๐ฆ_3`$, so that it is non-special.
Let $`0<\eta _1<ฯต`$. As above, we can then construct a non-special $`\alpha `$-limit point $`x_1`$ on the Kasner circle with $`\mathrm{\Sigma }_+`$ coordinate $`\mathrm{\Sigma }_{+,1}`$ such that $`1+\mathrm{\Sigma }_{+,1}\eta _1`$. Assume we have constructed non-special $`\alpha `$-limit points $`x_i`$ on the Kasner circle, $`i=1,\mathrm{},m`$ with $`\mathrm{\Sigma }_+`$ coordinates $`\mathrm{\Sigma }_{+,i}`$ satisfying $`\mathrm{\Sigma }_{+,i}<\mathrm{\Sigma }_{+,i1}`$. Let $`0<\eta _{m+1}<1+\mathrm{\Sigma }_{+,m}`$. Then by the above we can construct a non-special $`\alpha `$-limit point $`x_{m+1}`$ on the Kasner circle with $`\mathrm{\Sigma }_+`$ coordinate $`\mathrm{\Sigma }_{+,m+1}`$, satisfying $`\mathrm{\Sigma }_{+,m+1}<\mathrm{\Sigma }_{+,m}`$. Thus the solution has an infinite number of $`\alpha `$-limit points on the Kasner circle. $`\mathrm{}`$
###### Corollary 18.1.
A generic Bianchi IX solution with $`2/3<\gamma <2`$ has at least three non-special $`\alpha `$-limit points on the Kasner circle. Furthermore, no $`N_i`$ converges to zero.
Proof. Assume first that the solution has a special $`\alpha `$-limit point on the Kasner circle. By Lemma 18.1, the first part of the lemma follows. By the proof of Lemma 18.1, there is a non-special $`\alpha `$-limit point on the Kasner circle with $`\mathrm{\Sigma }_+`$ coordinate arbitrarily close to $`1`$, say that it belongs to $`๐ฆ_2`$. Repeated application of Proposition 6.1 then gives $`\alpha `$-limit points first in $`๐ฆ_3`$, and after enough iterates, either an $`\alpha `$-limit point in $`๐ฆ_1`$, or a special $`\alpha `$-limit point on the Kasner circle with $`\mathrm{\Sigma }_+=1/2`$. If the latter case occurs, a similar argument to the proof of Lemma 18.1 yields an $`\alpha `$-limit point on $`๐ฆ_1`$. By Proposition 6.1, we conclude that there are $`\alpha `$-limit points with $`N_1>0`$, with $`N_2>0`$ and with $`N_3>0`$.
Assume that there is no special $`\alpha `$-limit point on the Kasner circle. Repeated application of the Kasner map yields $`\alpha `$-limit points in $`๐ฆ_i`$, $`i=1,2,3`$, and the conclusions of the lemma follow as in the previous situation. $`\mathrm{}`$
## 19. Conclusions
Let us first state the conclusions concerning the asymptotics of solutions to the equations of Wainwright and Hsu. We begin with the stiff fluid case.
###### Theorem 19.1.
Consider a solution to (2.1)-(2.3) with $`\gamma =2`$ and $`\mathrm{\Omega }>0`$. Then the solution converges to a type I point with $`\mathrm{\Sigma }_+^2+\mathrm{\Sigma }_{}^2<1`$. For the Bianchi types other than I, we have the following additional restrictions.
1. If the solution is of type II with $`N_1>0`$, then $`\mathrm{\Sigma }_+<1/2`$.
2. For a type V$`I_0`$ or VI$`I_0`$ with $`N_2`$ and $`N_3`$ non-zero, then $`\mathrm{\Sigma }_+\pm \sqrt{3}\mathrm{\Sigma }_{}>1`$.
3. If the solution is of type VIII or IX, then $`\mathrm{\Sigma }_+\pm \sqrt{3}\mathrm{\Sigma }_{}>1`$ and $`\mathrm{\Sigma }_+<1/2`$.
Remark. Figure 8 illustrates the restriction on the shear variables. The types depicted are I, II, V$`\mathrm{I}_0`$ and VI$`\mathrm{I}_0`$, and VIII and IX, counting from top left to bottom right.
Proof. The theorem follows from Propositions 7.1 and 7.2. $`\mathrm{}`$
Consider now the case $`2/3<\gamma <2`$. Let $`๐`$ be the closure of the type II vacuum points.
###### Theorem 19.2.
Consider a generic Bianchi IX solution $`x`$ with $`2/3<\gamma <2`$. Then it converges to the closure of the set of vacuum type II points, that is
$$\underset{\tau \mathrm{}}{lim}\underset{y๐}{inf}x(\tau )y=0$$
where $``$ is the Euclidean norm on $`^6`$. Furthermore, there are at least three non-special $`\alpha `$-limit points on the Kasner circle.
Remark. One can start out arbitrarily close to this set without converging to it, cf. Proposition 11.1.
Proof. The first part follows from Corollary 15.4 and the second part follows from Corollary 18.1. $`\mathrm{}`$
Proof of Theorems 1.1 and 1.2. Let $`(M,g)`$ be the Lorentz manifold obtained in Lemma 21.2 with topology $`I\times G`$. It is globally hyperbolic by Lemma 21.4.
If the initial data satisfy $`\mathrm{tr}_gk=0`$ for a development not of type IX, then it is causally geodesically complete and satisfies $`\mu =0`$ for the entire development, by Lemma 21.5 and Lemma 21.8. The first part of Theorem 1.1 follows.
Consider initial data of type I, II, V$`\mathrm{I}_0`$, VI$`\mathrm{I}_0`$ or VIII such that $`\mathrm{tr}_gk0`$. By Lemma 21.5 and Lemma 21.8, we may then time orient the development so that it is future causally geodesically complete and past causally geodesically incomplete, and the second part of Theorem 1.1 follows. The third part follows from Lemma 21.8.
Consider an inextendible future directed causal geodesic in the above development. Since each hypersurface $`\{v\}\times G`$ is a Cauchy hypersurface by Lemma 21.4, the causal curve exhausts the interval $`I`$.
1. If the solution is not of type IX, then the solution to (21.4)-(21.9), which is used in constructing the class A development, corresponds to a solution to (2.1)-(2.3), because of Lemma 21.5. Furthermore, $`tt_{}`$ corresponds to $`\tau \mathrm{}`$, because of Lemma 22.4.
a. In all the stiff fluid cases, the solution to (2.1)-(2.3) converges to a non-vacuum type I point by Theorem 19.1, so that Lemma 22.1 and Lemma 22.3 yield the desired conclusions in that case.
b. Type I, II and VI$`\mathrm{I}_0`$ with $`1\gamma <2`$. That the Kretschmann scalar is unbounded in the cases stated in Theorem 1.2 follows from Proposition 8.1, Proposition 9.1, Proposition 10.2, Lemma 22.1 and and Lemma 22.2.
c. Non-vacuum solutions which are not of type IX. Then $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$ is unbounded using Lemma 22.3.
2. If the solution is of type IX, then half of a solution to (21.4)-(21.9) corresponds to a Bianchi IX solution to (2.1)-(2.3), because of Lemma 21.6. By Lemma 22.5, $`tt_\pm `$ corresponds to $`\tau \mathrm{}`$.
a. In the stiff fluid case, we get the desired statement as before.
b. If $`1\gamma <2`$, we get the desired conclusions, concerning blow up of the Kretschmann scalar, from Corollary 18.1, Proposition 11.1, Lemma 22.1 and Lemma 22.2.
c. Non-vacuum solutions. Then $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$ is unbounded using Lemma 22.3.
Let us now prove that the development is inextendible in the relevant cases. Assume there is a connected Lorentz manifold $`(\widehat{M},\widehat{g})`$ of the same dimension, and a map $`i:M\widehat{M}`$ which is an isometry onto its image, with $`i(M)\widehat{M}`$. Then there is a $`p\widehat{M}i(M)`$ and a timelike geodesic $`\gamma :[a,b]\widehat{M}`$ such that $`\gamma ([a,b))i(M)`$ and $`\gamma (b)=p`$. Since $`\gamma |_{[a,b)}`$ can be considered to be a future or past inextendible timelike geodesic in $`M`$, either it has infinite length or a curvature invariant blows up along it, by the above arguments. Both possibilities lead to a contradiction. Theorem 1.2 follows. $`\mathrm{}`$
## 20. Asymptotically velocity term dominated behaviour near the singularity
In this section, we consider the asymptotic behaviour of Bianchi VIII and IX stiff fluid solutions from another point of view. We wish to compare our results with , a paper which deals with analytic solutions of Einsteinโs equations coupled to a scalar field or a stiff fluid. In , Andersson and Rendall prove that given a certain kind of solution to the so called velocity dominated system, there is a unique solution of Einsteinโs equations coupled to a stiff fluid approaching the velocity dominated solution asymptotically. We will be more specific concerning the details below. The question which arises is to what extent it is natural to assume that a solution has the asymptotic behaviour they prescribe. We show here that all Bianchi VIII and IX stiff fluid solutions exhibit such asymptotic behaviour.
In order to speak about velocity term dominance, we need to have a foliation. In our case, there is a natural foliation given by the spatial hypersurfaces of homogeneity. Relative to this foliation, we can express the metric as in (21.14) according to Lemma 21.2. In what follows, we will use the frame $`e_i^{}`$ appearing in Lemma 21.2, and Latin indices will refer to this frame. Let $`g`$ be the Riemannian metric, and $`k`$ the second fundamental form of the spatial hypersurfaces of homogeneity, so that
(20.1)
$$g_{ij}=\overline{g}(e_i^{},e_j^{})=a_i^2\delta _{ij},$$
where $`\overline{g}`$ is as in (21.14). The constraint equations in our situation are
(20.2) $`Rk_{ij}k^{ij}+(\mathrm{tr}k)^2`$ $`=`$ $`2\mu `$
(20.3) $`^ik_{ij}_j(\mathrm{tr}k)`$ $`=`$ $`0,`$
which are the same as (21.8) and (21.5) respectively. The evolution equations are
(20.4) $`_tg_{ij}`$ $`=`$ $`2k_{ij}`$
(20.5) $`_tk_j^i`$ $`=`$ $`R_j^i+(\mathrm{tr}k)k_j^i.`$
The evolution equation for the matter is
(20.6)
$$_t\mu =2(\mathrm{tr}k)\mu .$$
We wish to compare solutions to these equations with solutions to the so called velocity dominated system. This system also consists of constraints and evolution equations, and we will denote the velocity dominated solution with a left superscript zero. The constraints are
(20.7) $`{}_{}{}^{0}k_{ij}^{}{}_{}{}^{0}k_{}^{ij}+(\mathrm{tr}{}_{}{}^{0}k)^2`$ $`=`$ $`2{}_{}{}^{0}\mu `$
(20.8) $`{}_{}{}^{0}_{}^{i}({}_{}{}^{0}k_{ij}^{}){}_{}{}^{0}_{j}^{}(\mathrm{tr}{}_{}{}^{0}k)`$ $`=`$ $`0.`$
The evolution equations are
(20.9) $`_t{}_{}{}^{0}g_{ij}^{}`$ $`=`$ $`2{}_{}{}^{0}k_{ij}^{}`$
(20.10) $`_t{}_{}{}^{0}k_{j}^{i}`$ $`=`$ $`(\mathrm{tr}{}_{}{}^{0}k){}_{}{}^{0}k_{j}^{i},`$
and the matter equation is
(20.11)
$$_t{}_{}{}^{0}\mu =2(\mathrm{tr}{}_{}{}^{0}k){}_{}{}^{0}\mu .$$
We raise and lower indices of the velocity dominated system with the velocity dominated metric. In , Andersson and Rendall prove that given an analytic solution to (20.7)-(20.11) on $`S\times (0,\mathrm{})`$ such that $`t\mathrm{tr}{}_{}{}^{0}k=1`$, and such that the eigenvalues of $`t{}_{}{}^{0}k_{j}^{i}`$ are positive, there is a unique analytic solution to (20.2)-(20.6) asymptotic, in a suitable sense, to the solution of the velocity dominated system. In fact, they prove this statement in a more general setting than the one given above. We have specialized to our situation. Observe the condition on the eigenvalues of $`t{}_{}{}^{0}k_{j}^{i}`$. Our goal is to prove that this is a natural condition in the Bianchi VIII and IX cases.
###### Theorem 20.1.
Consider a Bianchi VIII or IX stiff fluid development as in Lemma 21.2 with $`\mu _0>0`$. Choose time coordinate so that $`t_{}=0`$. Then there is a solution to (20.7)-(20.11) such that $`t\mathrm{tr}{}_{}{}^{0}k=1`$, the eigenvalues of $`t{}_{}{}^{0}k_{j}^{i}`$ are positive, and the following estimates hold
1. $`{}_{}{}^{0}g_{}^{il}g_{lj}=\delta _j^i+o(t^{\alpha _j^i})`$
2. $`k_j^i={}_{}{}^{0}k_{j}^{i}+o(t^{1+\alpha _j^i})`$
3. $`\mu ={}_{}{}^{0}\mu +o(t^{2+\beta _1})`$,
where $`\alpha _j^i`$ and $`\beta _1`$ are positive real numbers.
Remark. In two more estimates occur. They are not included here as they are replaced by equalities in our situation. Observe that the difficulties encountered in concerning the non-diagonal terms of $`k_j^i`$ disappear in the present situation.
Proof. Below we will use the results of Lemma 21.2 and its proof implicitly. When we speak of $`\theta _{ij}`$, $`\sigma _{ij}`$, $`\theta `$, $`n_{ij}`$ and $`\mu `$, we will refer to the solution of (21.4)-(21.9) and the indices of these objects should not be understood in terms of evaluation on a frame. Since $`\theta _{ij}`$ and so on are all diagonal, we will sometimes write $`\theta _i`$ etc instead, denoting diagonal component $`i`$. There are two relevant frames: $`e_i^{}`$ and $`e_i=a_ie_i^{}`$. The latter frame yields $`n_{ij}`$ through (1.7). When we speak of $`k_j^i`$, $`R_{ij}`$ and so on, we will always refer to the frame $`e_i^{}`$. We have
$$k_j^i=\theta _i\delta _j^i$$
(no summation on $`i`$). The metric is given by (20.1) above. Let us choose
(20.12)
$${}_{}{}^{0}k_{j}^{i}={}_{}{}^{0}\theta _{i}^{}\delta _j^i,$$
let $`{}_{}{}^{0}\theta ={}_{}{}^{0}\theta _{1}^{}+{}_{}{}^{0}\theta _{2}^{}+{}_{}{}^{0}\theta _{3}^{}`$ and
$${}_{}{}^{0}g_{ij}^{}={}_{}{}^{0}a_{i}^{2}\delta _{ij}$$
(no summation on $`i`$). Because of (20.12), equation (20.8) will be satisfied since it is a statement concerning the commutation of $`{}_{}{}^{0}k_{j}^{i}`$ and $`n_{ij}`$. The existence interval for the solution to Einsteinโs equations is $`(0,t_+)`$ by our conventions, and since we wish to have $`t\mathrm{tr}{}_{}{}^{0}k=1`$ we need to define $`{}_{}{}^{0}\theta (t)=1/t`$. Observe that $`{}_{}{}^{0}\theta _{i}^{}/{}_{}{}^{0}\theta `$ is constant in time, and that $`\theta _i/\theta `$ converges to a positive value as $`t0`$; this is a consequence of Theorem 19.1 and the definition (21.11) of the variables $`\mathrm{\Sigma }_+`$ and $`\mathrm{\Sigma }_{}`$. Choose $`{}_{}{}^{0}\theta _{i}^{}`$ so that $`{}_{}{}^{0}\theta _{i}^{}/{}_{}{}^{0}\theta `$ coincides with the limit of $`\theta _i/\theta `$. Similarly $`{}_{}{}^{0}\mu /{}_{}{}^{0}\theta _{}^{2}`$ is constant, $`\mu /\theta ^2`$ converges to a positive value, and we choose $`{}_{}{}^{0}\mu /{}_{}{}^{0}\theta _{}^{2}`$ to be the limit. Since $`R/\theta ^2`$ is a polynomial in the $`N_i`$ and the $`N_i`$ converge to zero by Theorem 19.1, equation (20.7) will be fulfilled. By our choices, (20.10) and (20.11) will also be fulfilled. We will specify the initial value of $`{}_{}{}^{0}a_{i}^{}`$ later on, and then define $`{}_{}{}^{0}a_{i}^{}`$ by demanding that (20.9) holds.
It will be of interest to estimate terms of the form $`R_j^i/\theta ^2`$. These terms are quadratic polynomials in the $`N_i`$. By abuse of notation, we will write $`N_i(\tau )`$ when we wish to evaluate $`N_i`$ in the Wainwright-Hsu time (21.10) and $`N_i(t)`$ when we wish to evaluate in the time used in this theorem. By Theorem 19.1, there is an $`ฯต>0`$ and a $`\tau _0`$ such that
$$|N_i(\tau )|\mathrm{exp}(ฯต\tau )$$
for all $`\tau \tau _0`$. We wish to rewrite this estimate in terms of $`t`$. Let us begin with (21.12). Since we can assume that $`q3`$ for $`\tau \tau _0`$ we get
$$\theta (\tau )\mathrm{exp}[4(\tau \tau _0)]\theta (\tau _0),$$
so that for $`\tau _1,\tau \tau _0`$ we get, using (21.10),
$$t(\tau )t(\tau _1)=_{\tau _1}^\tau \frac{3}{\theta }๐s\frac{3}{4\theta (\tau _0)}(\mathrm{exp}[4(\tau \tau _0)]\mathrm{exp}[4(\tau _1\tau _0)]).$$
Letting $`\tau _1`$ go to $`\mathrm{}`$ and observing that $`t(\mathrm{})=0`$, cf. Lemma 22.4 and Lemma 22.5, we get for some constant $`c`$
$$e^{4\tau }ct(\tau ),$$
so that
$$N_i(t)\mathrm{exp}(ฯต\tau (t))Ct^\eta $$
for some positive number $`\eta `$. Consequently expressions such as $`R_j^i/\theta ^2`$ and $`R/\theta ^2`$ satisfy similar bounds.
Let us now prove the estimates formulated in the statement of the theorem. Observe that for $`t`$ small enough, we have
$$\theta =\mathrm{tr}k(t)=(_0^t[\frac{R}{\theta ^2}+1]๐s)^1,$$
since the singularity is at $`t=0`$ and $`\mathrm{tr}k`$ must become unbounded at the singularity, cf. Lemma 22.4, 22.5 and (21.12). Thus we get
(20.13)
$$\theta {}_{}{}^{0}\theta =_0^t\frac{R}{\theta ^2}๐s\{t_0^t[\frac{R}{\theta ^2}+1]๐s\}^1=o(t^{1+\eta _1})$$
for some $`\eta _1>0`$. In order to make the estimates concerning $`k_j^i`$, we need only consider $`\theta _i`$ and $`{}_{}{}^{0}\theta _{i}^{}`$. We have
$$_t(\frac{\theta _i}{\theta }\frac{{}_{}{}^{0}\theta _{i}^{}}{{}_{}{}^{0}\theta })=_t\frac{\theta _i}{\theta }=\frac{\theta _iRR_i^i\theta }{\theta ^2}$$
with no summation on the $`i`$ in $`R_i^i`$. This computation, together with the estimates above and the fact that $`\theta _i/\theta {}_{}{}^{0}\theta _{i}^{}/{}_{}{}^{0}\theta `$ converges to zero, yields the estimate
(20.14)
$$\frac{\theta _i}{\theta }\frac{{}_{}{}^{0}\theta _{i}^{}}{{}_{}{}^{0}\theta }=o(t^{\eta _2}),$$
for some $`\eta _2>0`$. However,
(20.15)
$$\frac{\theta _i}{\theta }\frac{{}_{}{}^{0}\theta _{i}^{}}{{}_{}{}^{0}\theta }=\frac{\theta _i{}_{}{}^{0}\theta _{i}^{}}{\theta }+\frac{{}_{}{}^{0}\theta _{i}^{}}{{}_{}{}^{0}\theta }\frac{{}_{}{}^{0}\theta \theta }{\theta }.$$
Combining (20.13), (20.14) and (20.15), we get estimate $`2`$ of the theorem. Similarly, we have
$$_t(\frac{\mu }{\theta ^2}\frac{{}_{}{}^{0}\mu }{{}_{}{}^{0}\theta _{}^{2}})=_t\frac{\mu }{\theta ^2}=\frac{2\mu R}{\theta ^3}.$$
Integrating, using the fact that $`\mu /\theta ^2`$ converges to $`{}_{}{}^{0}\mu /{}_{}{}^{0}\theta _{}^{2}`$, we get
(20.16)
$$\frac{\mu }{\theta ^2}\frac{{}_{}{}^{0}\mu }{{}_{}{}^{0}\theta _{}^{2}}=o(t^{\eta _3})$$
where $`\eta _3>0`$. Using
$$\frac{\mu }{\theta ^2}\frac{{}_{}{}^{0}\mu }{{}_{}{}^{0}\theta _{}^{2}}=\frac{\mu {}_{}{}^{0}\mu }{\theta ^2}+\frac{{}_{}{}^{0}\mu }{{}_{}{}^{0}\theta _{}^{2}}\frac{{}_{}{}^{0}\theta _{}^{2}\theta ^2}{\theta ^2},$$
(20.13) and (20.16), we get estimate $`3`$ of the theorem. Finally, we need to specify the initial value of $`{}_{}{}^{0}a_{i}^{}`$ and prove estimate $`1`$. Since
$$_ta_i=\theta _ia_i,$$
(no summation on $`i`$) and similarly for $`{}_{}{}^{0}a_{i}^{}`$, we get
$$_t\frac{a_i}{{}_{}{}^{0}a_{i}^{}}=\frac{a_i}{{}_{}{}^{0}a_{i}^{}}({}_{}{}^{0}\theta _{i}^{}\theta _i).$$
By our estimates on $`{}_{}{}^{0}\theta _{i}^{}\theta _i`$, we see that this implies that $`a_i/{}_{}{}^{0}a_{i}^{}`$ converges as $`t0`$. Choose the value of $`{}_{}{}^{0}a_{i}^{}`$ at one point in time so that this limit is $`1`$. We thus get, using estimate $`2`$ of the theorem,
$$\frac{a_i}{{}_{}{}^{0}a_{i}^{}}1=o(t^{\alpha _j^i}).$$
Estimate $`1`$ of the theorem now follows from this estimate and the fact that
$${}_{}{}^{0}g_{}^{il}g_{lj}=\left(\frac{{}_{}{}^{0}a_{i}^{}}{a_i}\right)^2\delta _j^i.$$
The theorem follows. $`\mathrm{}`$
## 21. Appendix
The goal of this appendix is to relate the asymptotic behaviour of solutions to the ODE (2.1)-(2.3) to the behaviour of the spacetime in the incomplete directions of inextendible causal curves. We proceed as follows.
1. First, we formulate Einsteinโs equations as an ODE, assuming that the spacetime has a given structure (21.1). The first formulation is due to Ellis and MacCallum. We also relate this formulation to the one by Wainwight and Hsu.
2. Given initial data as in Definition 1.1, we then show how to construct a Lorentz manifold as in (21.1), satisfying Einsteinโs equations and with initial data as specified, using the equations of Ellis and MacCallum. We also prove some properties of this development such as Global hyperbolicity and answer some questions concerning causal geodesic completeness.
3. Finally, we relate the asymptotic behaviour of solutions to (2.1)-(2.3) to the question of curvature blow up in the development obtained by the above procedure.
We consider a special class of spatially homogeneous four dimensional spacetimes of the form
(21.1)
$$(\overline{M},\overline{g})=(I\times G,dt^2+\chi _{ij}(t)\xi ^i\xi ^j),$$
where $`I`$ is an open interval, $`G`$ is a Lie group of class A, $`\chi _{ij}`$ is a smooth positive definite matrix and the $`\xi ^i`$ are the duals of a left invariant basis on $`G`$. The stress energy tensor is assumed to be given by
(21.2)
$$T=\mu dt^2+p(\overline{g}+dt^2),$$
where $`p=(\gamma 1)\mu `$. Below, Latin indices will be raised and lowered by $`\delta _{ij}`$.
Consider a four dimensional $`(\overline{M},\overline{g})`$ as in (21.1) with $`G`$ of class A. In order to define the different variables, we specify a suitable orthonormal basis. Let $`e_0=_t`$ and $`e_i=a_i^jZ_j`$, i=1,2,3, be an orthonormal basis, where $`a`$ is a $`C^{\mathrm{}}`$ matrix valued function of $`t`$ and the $`Z_i`$ are the duals of $`\xi ^i`$.
By the following argument, we can assume that $`<\overline{}_{e_0}e_i,e_j>=0`$. Let the matrix valued function $`A`$ satisfy $`e_0(A)+AB=0`$, $`A(0)=\mathrm{Id}`$ where $`B_{ij}=<\overline{}_{e_0}e_i,e_j>`$ and $`\mathrm{Id}`$ is the $`3\times 3`$ identity matrix. Then $`A`$ is smooth and $`SO(3)`$ valued and if $`e_i^{}=A_i^je_j`$, then $`<\overline{}_{e_0}e_i^{},e_j^{}>=0`$.
Let
(21.3)
$$\theta (X,Y)=<\overline{}_Xe_0,Y>,$$
$`\theta _{\alpha \beta }=\theta (e_\alpha ,e_\beta )`$ and $`[e_\beta ,e_\gamma ]=\gamma _{\beta \gamma }^\alpha e_\alpha `$ where Greek indices run from $`0`$ to $`3`$. The objects $`\theta _{\alpha \beta }`$ and $`\gamma _{\beta \gamma }^\alpha `$ will be viewed as smooth functions from $`I`$ to some suitable $`^k`$, and our variables will be defined in terms of them.
Observe that $`[Z_i,e_0]=0`$. The $`e_i`$ span the tangent space of $`G`$, and $`<[e_0,e_i],e_0>=0`$. We get $`\theta _{00}=\theta _{0i}=0`$ and $`\theta _{\alpha \beta }`$ symmetric. We also have $`\gamma _{ij}^0=\gamma _{0i}^0=0`$ and $`\gamma _{0j}^i=\theta _{ij}`$. We let $`n`$ be defined as in (1.7) and
$$\sigma _{ij}=\theta _{ij}\frac{1}{3}\theta \delta _{ij},$$
where we by abuse of notation have written $`\mathrm{tr}(\theta )`$ as $`\theta `$.
We express Einsteinโs equations in terms of $`n`$, $`\sigma `$ and $`\theta `$. The Jacobi identities for $`e_\alpha `$ yield
(21.4)
$$e_0(n_{ij})2n_{k(i}\sigma _{j)}^k+\frac{1}{3}\theta n_{ij}=0.$$
The $`0i`$-components of the Einstein equations are equivalent to
(21.5)
$$\sigma _i^kn_{kj}n_i^k\sigma _{kj}=0.$$
Letting $`b_{ij}=2n_i^kn_{kj}\mathrm{tr}(n)n_{ij}`$ and $`s_{ij}=b_{ij}\frac{1}{3}\mathrm{tr}(b)\delta _{ij}`$, the trace free part of the $`ij`$ equations are
(21.6)
$$e_0(\sigma _{ij})+\theta \sigma _{ij}+s_{ij}=0.$$
The $`00`$-component yields the Raychaudhuri equation
(21.7)
$$e_0(\theta )+\theta _{ij}\theta ^{ij}+\frac{1}{2}(3\gamma 2)\mu =0,$$
and using this together with the trace of the $`ij`$-equations yields a constraint
(21.8)
$$\sigma _{ij}\sigma ^{ij}+(n_{ij}n^{ij}\frac{1}{2}\mathrm{tr}(n)^2)+2\mu =\frac{2}{3}\theta ^2.$$
Equations (21.4)-(21.8) are special cases of equations given in Ellis and MacCallum . At a point $`t_0`$, we may diagonalize $`n`$ and $`\sigma `$ simultaneously since they commute (21.5). Rotating $`e_\alpha `$ by the corresponding element of $`SO(3)`$ yields upon going through the definitions that the new $`n`$ and $`\sigma `$ are diagonal at $`t_0`$. Collect the off-diagonal terms of $`n`$ and $`\sigma `$ in one vector $`v`$. By (21.4) and (21.6), there is a time dependent matrix $`C`$ such that $`\dot{v}=Cv`$ so that $`v(t)=0`$ for all $`t`$, since $`v(t_0)=0`$. Since the rotation was time independent, $`<_{e_0}e_i,e_j>=0`$ holds in the new basis.
The fact that $`T`$ is divergence free yields
(21.9)
$$e_0(\mu )+\gamma \theta \mu =0.$$
Introduce, as in Wainwright and Hsu ,
$`\mathrm{\Sigma }_{ij}=\sigma _{ij}/\theta `$
$`N_{ij}=n_{ij}/\theta `$
$`\mathrm{\Omega }=3\mu /\theta ^2`$
and define a new time coordinate $`\tau `$, independent of time orientation, satisfying
(21.10)
$$\frac{dt}{d\tau }=\frac{3}{\theta }.$$
For Bianchi IX developments, we only consider the part of spacetime where $`\theta `$ is strictly positive or strictly negative. Let
(21.11)
$$\mathrm{\Sigma }_+=\frac{3}{2}(\mathrm{\Sigma }_{22}+\mathrm{\Sigma }_{33})\mathrm{and}\mathrm{\Sigma }_{}=\frac{\sqrt{3}}{2}(\mathrm{\Sigma }_{22}\mathrm{\Sigma }_{33}).$$
If we let $`N_i`$ be the diagonal elements of $`N_{ij}`$, equations (21.4) and (21.6) turn into (2.1) with definitions as in (2.2), except for the expression for $`\mathrm{\Omega }^{}`$. It can however be derived from (21.9). The constraint (21.8) turns into (2.3). The Raychaudhuri equation (21.7) takes the form
(21.12)
$$\theta ^{}=(1+q)\theta .$$
Before using the equations of Ellis and MacCallum to construct a development, it is convenient to know that one can make some simplifying assumptions concerning the choice of basis. The next lemma fulfills this objective, and also proves the classification of the class A Lie algebras mentioned in the introduction.
###### Lemma 21.1.
Table 1 constitutes a classification of the class A Lie algebras. Consider an arbitrary basis $`\{e_i\}`$ of the Lie algebra. Then by applying an orthogonal matrix to it, we can construct a basis $`\{e_i^{}\}`$ such that the corresponding $`n^{}`$ defined by (1.7) is diagonal, with diagonal elements of one of the types given in Table 1.
Proof. Let $`e_i`$ be a basis for the Lie algebra and $`n`$ be defined as in (1.7). If we change the basis according to $`e_i^{}=(A^1)_i^je_j`$, then $`n`$ transforms to
(21.13)
$$n^{}=(detA)^1A^tnA$$
Since $`n`$ is symmetric, we assume from here on that the basis is such that it is diagonal. The matrix $`A=\mathrm{diag}(\mathrm{1\; 1}1)`$ changes the sign of $`n`$. A suitable orthogonal matrix performs even permutations of the diagonal. The number of non-zero elements on the diagonal is invariant under transformations (21.13) taking one diagonal matrix to another. If $`A=(a_{ij})`$ and the diagonal matrix $`n^{}`$ is constructed as in (21.13), we have $`n_{kk}^{}=(detA)^1_{i=1}^3a_{ik}^2n_{ii}`$, so that if all the diagonal elements of $`n`$ have the same sign, the same is true for $`n^{}`$. The statements of the lemma follow. $`\mathrm{}`$
We now prove that if we begin with initial data as in Definition 1.1, we get a development as in Definition 1.4 of the form (21.1), with certain properties.
###### Lemma 21.2.
Fix $`2/3<\gamma 2`$. Let $`G,g,k`$ and $`\mu _0`$ be initial data as in Definition 1.1. Then there is an orthonormal basis $`e_i^{}`$ $`i=1,2,3`$ of the Lie algebra such that $`n_{ij}^{}`$ defined by (1.7) and $`k_{ij}=k(e_i^{},e_j^{})`$ are diagonal and $`n_{ij}^{}`$ is of one of the forms given in Table 1. Let
$$\theta (0)=\mathrm{tr}_gk,\sigma _{ij}(0)=k(e_i^{},e_j^{})+\frac{1}{3}\theta (0)\delta _{ij},n_{ij}(0)=n_{ij}^{}and\mu (0)=\mu _0.$$
Solve (21.4), (21.6), (21.7) and (21.9) with these conditions as initial data to obtain $`n,\sigma ,\theta `$ and $`\mu `$, and let $`I`$ be the corresponding existence interval. Then there are smooth functions $`a_i:I(0,\mathrm{})`$ $`i=1,2,3`$, with $`a_i(0)=1`$, such that
(21.14)
$$\overline{g}=dt^2+\underset{i=1}{\overset{3}{}}a_i^2(t)\xi ^i\xi ^i,$$
where $`\xi ^i`$ is the dual of $`e_i^{}`$, satisfies Einsteinโs equations (1.3) on $`\overline{M}=I\times G`$, with $`T`$ as in (1.1) with $`u=e_0`$, $`\mu `$ as above and $`p=(\gamma 1)\mu `$. Furthermore,
$$<\overline{}_{e_i}e_0,e_j>=\sigma _{ij}+\frac{1}{3}\theta \delta _{ij},$$
where $`\overline{}`$ is the Levi-Civita connection of $`\overline{g}`$ and $`e_i=a_ie_i^{}`$, if we consider the left hand side to be a function of $`t`$. Consequently, the induced metric and second fundamental form on $`\{0\}\times G`$ are $`g`$ and $`k`$, and we have a development satisfying the conditions of Definition 1.4.
Proof. Let $`e_i^{}`$, $`i=1,2,3`$ be a left invariant orthonormal basis. We can assume the corresponding $`n^{}`$ to be of one of the forms given in Table 1 by Lemma 21.1. The content of (1.5) is that $`k_{ij}=k(e_i^{},e_j^{})`$ and $`n^{}`$ are to commute. We may thus also assume $`k_{ij}`$ to be diagonal without changing the earlier conditions of the construction. If we let $`n(0)=n^{}`$, $`\theta (0)=\mathrm{tr}_gk`$, $`\sigma _{ij}(0)=k_{ij}+\theta \delta _{ij}/3`$ and $`\mu (0)=\mu _0`$, then (1.4) is the same as (21.8). Let $`n`$, $`\sigma `$, $`\theta `$ and $`\mu `$ satisfy (21.4), (21.6), (21.7) and (21.9) with initial values as specified above. Since (21.8) is satisfied at $`0`$, it is satisfied for all times. For reasons given in connection with (21.8), $`n`$ and $`\sigma `$ will remain diagonal so that (21.5) will always hold. Let $`n_i`$ and $`\sigma _i`$ denote the diagonal elements of $`n`$ and $`\sigma `$ respectively.
How are we to define the $`a_i`$ in the statement of the lemma? The $`\stackrel{~}{n}`$ obtained from $`e_i`$ by (1.7) should coincide with $`n`$. This leads us to the following definitions. Let $`f_i(0)=1`$ and $`\dot{f_i}/f_i=2\sigma _i\theta /3`$. Let $`a_i=(\mathrm{\Pi }_{ji}f_j)^{1/2}`$ and define $`e_i=a_ie_i^{}`$. Then $`\stackrel{~}{n}`$ associated to $`e_i`$ equals $`n`$. We complete the basis by letting $`e_0=_t`$. Define a metric $`<,>`$ on $`\overline{M}`$ by demanding $`e_\alpha `$ to be orthonormal with $`e_0`$ timelike and $`e_i`$ spacelike, and let $`\overline{}`$ be the associated Levi-Civita connection. Compute $`<\overline{}_{e_0}e_i,e_j>=0`$. If $`\stackrel{~}{\theta }(X,Y)=<\overline{}_Xe_0,Y>`$ and $`\stackrel{~}{\theta }_{\mu \nu }=\stackrel{~}{\theta }(e_\mu ,e_\nu )`$, then $`\stackrel{~}{\theta }_{00}=\stackrel{~}{\theta }_{i0}=\stackrel{~}{\theta }_{0i}=0`$. Furthermore,
$$\frac{1}{a_j}e_0(a_j)\delta _{ij}=\stackrel{~}{\theta }_{ij}$$
(no summation over $`j`$) so that $`\stackrel{~}{\theta }_{ij}`$ is diagonal and $`\mathrm{tr}\stackrel{~}{\theta }=\theta `$. Finally,
$$\stackrel{~}{\sigma }_{ii}=\stackrel{~}{\theta }_{ii}+\frac{1}{3}\theta =\sigma _i.$$
The lemma follows by considering the derivation of the equations of Ellis and MacCallum. $`\mathrm{}`$
###### Definition 21.1.
A development as in Lemma 21.2 will be called a class A development. We will also assign a type to such a development according to the type of the initial data.
The next thing to prove is that each $`\overline{M}_v=\{v\}\times G`$ is a Cauchy surface, but first we need a lemma.
###### Lemma 21.3.
Let $`\rho `$ be a left invariant Riemannian metric on a Lie group $`G`$. Then $`\rho `$ is geodesically complete.
Proof. Assume $`\gamma :(t_{},t_+)G`$ is a geodesic satisfying $`\rho (\gamma ^{},\gamma ^{})=1`$, with $`t_+<\mathrm{}`$. There is a $`\delta >0`$ such that every geodesic $`\lambda `$ satisfying $`\lambda (0)=e`$, the identity element of $`G`$, and $`\lambda ^{}(0)=v`$ with $`\rho (v,v)1`$ is defined on $`(\delta ,\delta )`$. If $`L_h:GG`$ is defined by $`L_h(h_1)=hh_1`$, then $`L_h`$ is by definition an isometry. Let $`t_0(t_{},t_+)`$ satisfy $`t_+t_0\delta /2`$. Let $`vT_eG`$ be the vector corresponding to $`\gamma ^{}(t_0)`$ under the isometry $`L_{\gamma (t_0)}`$. Let $`\lambda `$ be a geodesic with $`\lambda (0)=e`$ and $`\lambda ^{}(0)=v`$. Then $`L_{\gamma (t_0)}\lambda `$ is a geodesic extending $`\gamma `$. $`\mathrm{}`$
Let us be precise concerning the concept Cauchy surface.
###### Definition 21.2.
Consider a time oriented Lorentz manifold $`(M,g)`$. Let $`I`$ be an interval in $``$ and $`\gamma :IM`$ be a continuous map which is smooth except for a finite number of points. We say that $`\gamma `$ is a future directed causal, timelike or null curve if at each $`tI`$ where $`\gamma `$ is differentiable, $`\gamma ^{}(t)`$ is a future oriented causal, timelike or null vector respectively. We define past directed curves similarly. A causal curve is a curve which is either a future directed causal curve or a past directed causal curve and similarly for timelike and null curves. If there is a curve $`\lambda :I_1M`$ such that $`\gamma (I)`$ is properly contained in $`\lambda (I_1)`$, then $`\gamma `$ is said to be extendible, otherwise it is called inextendible. A subset $`SM`$ is called a Cauchy surface if it is intersected exactly once by every inextendible causal curve. A Lorentz manifold as above which admits a Cauchy surface is said to be Globally hyperbolic.
###### Lemma 21.4.
For a class A development, each $`\overline{M}_v=\{v\}\times G`$ is a Cauchy surface.
Proof. The metric is given by (21.14). A causal curve cannot intersect $`\overline{M}_v`$ twice since the $`t`$-component of such a curve must be strictly monotone. Assume that $`\gamma :(s_{},s_+)M`$ is an inextendible causal curve that never intersects $`\overline{M}_v`$. Let $`\stackrel{~}{t}:\overline{M}I`$ be defined by $`\stackrel{~}{t}[(s,h)]=s`$. Let $`s_0(s_{},s_+)`$ and assume that $`\stackrel{~}{t}(\gamma (s_0))=t_1<v`$ and that $`<\gamma ^{},_t><0`$ where it is defined. Thus $`\stackrel{~}{t}(\gamma (s))`$ increases with $`s`$ and $`\stackrel{~}{t}(\gamma ([s_0,s_+)))[t_1,v]`$. Since we have uniform bounds on $`a_i`$ from below and above on $`[t_1,v]`$ and the curve is causal, we get
(21.15)
$$(\underset{i=1}{\overset{3}{}}\xi ^i(\gamma ^{})^2)^{1/2}C<\gamma ^{},e_0>$$
on that interval, with $`C>0`$. Since
(21.16)
$$_{s_0}^{s_+}<\gamma ^{},e_0>ds=_{s_0}^{s_+}\frac{d\stackrel{~}{t}\gamma }{ds}๐svt_1,$$
the curve $`\gamma |_{[s_0,s_+)}`$, projected to $`G`$, will have finite length in the metric $`\rho `$ on $`G`$ defined by making $`e_i^{}`$ an orthonormal basis. Since $`\rho `$ is a left invariant metric on a Lie group, it is complete by Lemma 21.3, and sets closed and bounded in the corresponding topological metric must be compact. Adding the above observations, we conclude that $`\gamma ([s_0,s_+))`$ is contained in a compact set, and thus there is a sequence $`s_k[s_0,s_+)`$ with $`s_ks_+`$ such that $`\gamma (s_k)`$ converges. Since $`\stackrel{~}{t}(\gamma (s))`$ is monotone and bounded it converges. Using (21.15) and an analogue of (21.16), we conclude that $`\gamma `$ has to converge as $`ss_+`$. Consequently, $`\gamma `$ is extendible contradicting our assumption. By this and similar arguments covering the other cases, we conclude that $`\overline{M}_v`$ is a Cauchy surface for each $`v(t_{},t_+)`$. $`\mathrm{}`$
Before we turn to the questions concerning causal geodesic completeness, let us consider the evolution of $`\theta `$ for solutions to the equations of Ellis and MacCallum. This is relevant also for the definition of the variables of Wainwright and Hsu, since there one divides by $`\theta `$. We first consider developments as in Lemma 21.2 which are not of type IX.
###### Lemma 21.5.
Consider class A developments which are not of type IX. Let the existence interval be $`I=(t_{},t_+)`$. Then there are two possibilities.
1. $`\theta 0`$ for the entire development. We then time orient the manifold so that $`\theta >0`$. With this time orientation, $`t_+=\mathrm{}`$.
2. $`\theta =0`$, $`\sigma _{ij}=0`$ and $`\mu =0`$ for the entire development. Furthermore, $`n_{ij}`$ is constant and diagonal and two of the diagonal components are equal and the third is zero. The only Bianchi types which admit this possibility are thus type I and type VI$`I_0`$. Furthermore $`I=(\mathrm{},\mathrm{})`$.
Proof. Since $`n_{ij}`$ is diagonal, see the proof of Lemma 21.2, we can formulate the constraint (21.8) as
$$\sigma _{ij}\sigma ^{ij}+\frac{1}{2}[n_1^2+(n_2n_3)^22n_1(n_2+n_3)]+2\mu =\frac{2}{3}\theta ^2,$$
where the $`n_i`$ are the diagonal components of $`n_{ij}`$. Considering Table 1, we see that, excepting type IX, the expression in the $`n_i`$ is always non-negative. Thus we deduce the inequality
(21.17)
$$\sigma _{ij}\sigma ^{ij}+2\mu \frac{2}{3}\theta ^2.$$
Combining it with (21.7), we get $`|e_0(\theta )|\theta ^2`$, using the fact that $`2/3<\gamma 2`$. Consequently, if $`\theta `$ is zero once, it is always zero. Time orient the developments with $`\theta 0`$ so that $`\theta >0`$.
Consider the possibility $`\theta =0`$. Equation (21.7) then implies $`\sigma _{ij}=0`$ and $`\mu =0`$, since $`\gamma >2/3`$. Equations (21.8) and (21.6) then imply $`b_{ij}=0`$, and (21.4) implies $`n_{ij}`$ constant. All the statements except the the fact that $`t_+=\mathrm{}`$ in the $`\theta >0`$ case follow from the above.
Observe that $`\theta `$ decreases in magnitude with time, so that it is bounded to the future. By the (21.17), the same is true of $`\sigma _{ij}`$ and $`\mu `$. Using (21.4), we get control of $`n_{ij}`$ and conclude that the solution may not blow up in finite time. We must thus have $`t_+=\mathrm{}`$. $`\mathrm{}`$
By a theorem of Lin and Wald , Bianchi IX developments recollapse.
###### Lemma 21.6.
Consider a Bianchi IX class A development with $`1\gamma 2`$ and $`I=(t_{},t_+)`$. Then there is a $`t_0I`$ such that $`\theta >0`$ in $`(t_{},t_0)`$ and $`\theta <0`$ in $`(t_0,t_+)`$.
Proof. Let us begin by proving that $`\theta `$ can be zero at most once. If $`\theta (t_i)=0`$, $`i=1,2`$ and $`t_1<t_2`$, then $`\theta =0`$ in $`(t_1,t_2)`$ since it is monotone by (21.7). Thus (21.7) implies $`\sigma _{ij}=0=\mu `$ in $`(t_1,t_2)`$ as well. Combining this fact with (21.8) and (21.6), we get $`b_{ij}=0`$, which is impossible for a Bianchi IX solution. Assume $`\theta `$ is never zero. By a suitable choice of time orientation, we can assume that $`\theta >0`$ on $`I`$. Let us prove that $`t_+=\mathrm{}`$. Since $`\theta `$ is decreasing on $`I_1=[0,t_+)`$ and non-negative on $`I`$ it is bounded on $`I_1`$. By (21.4), $`n_1n_2n_3`$ decreases so that it is bounded on $`I_1`$. By an argument similar to the proof of Lemma 3.3, one can combine this bound with (21.8) to conclude that $`\sigma _{ij}`$ and $`\mu `$ are bounded on $`I_1`$. By (21.4), we conclude that $`n_{ij}`$ cannot grow faster than exponentially. Consequently, the future existence interval must be infinite, that is $`t_+=\mathrm{}`$, since $`I`$ was the maximal existence interval and solutions cannot blow up in finite time. In order to use the arguments of Lin and Wald, we define
$$\beta _i(t)=_0^t\sigma _i(s)๐s+\beta _i^0,\alpha (t)=_0^t\frac{1}{3}\theta (s)๐s+\alpha _0,$$
where $`2\beta _i^0\alpha _0=\mathrm{ln}(n_i(0))`$ and $`_{i=1}^3\beta _i^0=0`$. Then
$$n_i=\mathrm{exp}(2\beta _i\alpha ).$$
Let $`\rho =\mu /8\pi `$ and $`P_i=p/8\pi =(\gamma 1)\mu /8\pi `$, $`i=1,2,3`$. Equations (21.8) and (21.7) then imply equations (1.4) and (1.5) of , and equations (1.6) and (1.7) of follow from (21.6). We have thus constructed a solution to (1.4)-(1.7) of on an interval $`[0,\mathrm{})`$ with $`d\alpha /dt>0`$. Lin and Wald prove in their paper that this assumption leads to a contradiction, if one assumes that $`|P_i|\rho `$ and $`P_1+P_2+P_30`$. However, these conditions are fulfilled in our situation, assuming $`1\gamma 2`$. In other words, there is a zero and since $`\theta `$ is decreasing it must be positive before the zero and negative after it. The lemma follows. $`\mathrm{}`$
The lemma concerning causal geodesic completeness will build on the following estimate.
###### Lemma 21.7.
Consider a class A development. Let $`\gamma :(s_{},s_+)\overline{M}`$ be a future directed inextendible causal geodesic, and
(21.18)
$$f_\nu (s)=<\gamma ^{}(s),e_\nu |_{\gamma (s)}>.$$
If $`\theta =0`$ for the entire development, then $`f_0`$ is constant. Otherwise,
(21.19)
$$\frac{d}{ds}(f_0\theta )\frac{2\sqrt{2}}{3}\theta ^2f_0^2.$$
Remark. We consider functions of $`t`$ as functions of $`s`$ by evaluating them at $`\stackrel{~}{t}(\gamma (s))`$, where $`\stackrel{~}{t}`$ is the function defined in Lemma 21.4.
Proof. Compute, using the proof of Lemma 21.2,
$$\frac{df_0}{ds}=<\gamma ^{}(s),_{\gamma ^{}(s)}e_0>=\underset{k=1}{\overset{3}{}}\theta _kf_k^2,$$
where $`\theta _k`$ are the diagonal elements of $`\theta _{ij}`$. If $`\theta =0`$ for the entire development, then $`\theta _k=0`$ for the entire development by Lemma 21.5 and Lemma 21.6, so that $`f_0`$ is constant. Compute, using Raychaudhuriโs equation (21.7),
$$\frac{d}{ds}(f_0\theta )=\frac{1}{3}\theta ^2\underset{k=1}{\overset{3}{}}f_k^2+\underset{k=1}{\overset{3}{}}\theta \sigma _kf_k^2+f_0^2\underset{k=1}{\overset{3}{}}\sigma _k^2+\frac{1}{3}\theta ^2f_0^2+\frac{1}{2}(3\gamma 2)\mu f_0^2$$
where $`\sigma _k`$ are the diagonal elements of $`\sigma _{ij}`$. Estimate
$$|\underset{k=1}{\overset{3}{}}\sigma _kf_k^2|\left(\frac{2}{3}\right)^{1/2}\left(\underset{k=1}{\overset{3}{}}\sigma _k^2\right)^{1/2}\underset{k=1}{\overset{3}{}}f_k^2,$$
using the tracelessness of $`\sigma _{ij}`$. By making a division into the three cases $`_{k=1}^3\sigma _k^2\theta ^2/3`$, $`\theta ^2/3_{k=1}^3\sigma _k^22\theta ^2/3`$ and $`2\theta ^2/3_{k=1}^3\sigma _k^2`$, and using the causality of $`\gamma `$ we deduce (21.19). $`\mathrm{}`$
###### Lemma 21.8.
Consider a class A development with existence interval $`I=(t_{},t_+)`$. There are three possibilities.
1. $`\theta =0`$ for the entire development, in which case the development is causally geodesically complete.
2. The development is not of type IX and $`\theta >0`$. Then all inextendible causal geodesics are future complete and past incomplete. Furthermore, $`t_{}>\mathrm{}`$ and $`t_+=\mathrm{}`$.
3. If the development is of type IX with $`1\gamma 2`$, then all inextendible causal geodesics are past and future incomplete. We also have $`t_{}>\mathrm{}`$ and $`t_+<\mathrm{}`$.
Proof. Let $`\gamma :(s_{},s_+)\overline{M}`$ be a future directed inextendible causal geodesic and $`f_\nu `$ be defined as in (21.18). Let furthermore $`I=(t_{},t_+)`$ be the existence interval mentioned in Lemma 21.2. Since every $`\overline{M}_v`$, $`vI`$ is a Cauchy surface by Lemma 21.4, $`\stackrel{~}{t}(\gamma (s))`$ must cover the interval $`I`$ as $`s`$ runs through $`(s_{},s_+)`$. Furthermore, $`\stackrel{~}{t}(\gamma (s))`$ is monotone increasing so that
(21.20)
$$\stackrel{~}{t}(\gamma (s))t_\pm \mathrm{as}ss_\pm .$$
Let $`s_0(s_{},s_+)`$ and compute
(21.21)
$$_{s_0}^sf_0(u)du=\stackrel{~}{t}(\gamma (s))\stackrel{~}{t}(\gamma (s_0)).$$
Consider the case $`\theta =0`$ for the entire development. By Lemma 21.7, $`f_0`$ is then constant, and $`I=(\mathrm{},\mathrm{})`$ by Lemma 21.5. Equations (21.21) and (21.20) then prove that we must have $`(s_{},s_+)=(\mathrm{},\mathrm{})`$. Thus, all inextendible causal geodesics must be complete.
Assume that the development is not of type IX and that $`\theta >0`$. Since $`f_0\theta `$ is negative on $`[s_0,s_+)`$, its absolute value is bounded on that interval by (21.19). If $`s_+`$ were finite, $`\theta `$ would be bounded from below by a positive constant on $`[s_0,s_+)`$, since
$$|\frac{d\theta }{ds}|f_0\theta ^2C\theta $$
on that interval for some $`C>0`$, cf. (21.17) and the observations following that equation. Since $`f_0\theta `$ is bounded, we then deduce that $`f_0`$ is bounded on $`[s_0,s_+)`$. But then (21.20) and (21.21) cannot both hold, since $`t_+=\mathrm{}`$ by Lemma 21.5. Thus, $`s_+=\mathrm{}`$ and all inextendible causal geodesics must future complete. Since $`f_0\theta `$ is negative on $`(s_{},s_+)`$, (21.19) proves that this expression must blow up in finite $`s`$-time going backward, so that $`s_{}>\mathrm{}`$. Since the curve $`\gamma (s)=(s,e)`$ is an inextendible timelike geodesic, we conclude that $`t_{}>\mathrm{}`$.
Consider the Bianchi IX case. By Lemma 21.4 and 21.6, we conclude the existence of an $`s_0(s_{},s_+)`$ such that $`f_0\theta `$ is negative on $`(s_{},s_0)`$ and positive on $`(s_0,s_+)`$. By (21.19), $`f_0\theta `$ must blow up a finite $`s`$-time before $`s_0`$, and a finite $`s`$-time after $`s_0`$. Every inextendible causal geodesic is thus future and past incomplete. We conclude $`t_{}>\mathrm{}`$ and $`t_+<\mathrm{}`$. $`\mathrm{}`$
## 22. Appendix
In this appendix, we consider the curvature expressions. According to , p. 40, the Weyl tensor $`\overline{C}_{\alpha \beta \gamma \delta }`$ is defined by
$$\overline{R}_{\alpha \beta \gamma \delta }=\overline{C}_{\alpha \beta \gamma \delta }+(\overline{g}_{\alpha [\gamma }\overline{R}_{\delta ]\beta }\overline{g}_{\beta [\gamma }\overline{R}_{\delta ]\alpha })\frac{1}{3}\overline{R}\overline{g}_{\alpha [\gamma }\overline{g}_{\delta ]\beta },$$
where the bar in $`\overline{g}_{\alpha \beta }`$ and so on indicates that we are dealing with spacetime objects as opposed to objects on a spatial hypersurface. Using this relation and the fact that our spacetime satisfies (1.3), where $`T`$ is given by (1.1) and (1.2), one can derive the following expression for the Kretschmann scalar
(22.1)
$$\kappa =\overline{R}_{\alpha \beta \gamma \delta }\overline{R}^{\alpha \beta \gamma \delta }=\overline{C}_{\alpha \beta \gamma \delta }\overline{C}^{\alpha \beta \gamma \delta }+2\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }\frac{1}{3}\overline{R}^2=$$
$$=\overline{C}_{\alpha \beta \gamma \delta }\overline{C}^{\alpha \beta \gamma \delta }+\frac{1}{3}[4+(3\gamma 2)^2]\mu ^2.$$
However, according to , p. 19, we have
(22.2)
$$\overline{C}_{\alpha \beta \gamma \delta }\overline{C}^{\alpha \beta \gamma \delta }=8(E_{\alpha \beta }E^{\alpha \beta }H_{\alpha \beta }H^{\alpha \beta }),$$
where, relative to the frame $`e_\alpha `$ appearing in Lemma 21.2, all components of $`E`$ and $`H`$ involving $`e_0`$ are zero, and the $`ij`$ components are given by
$`E_{ij}`$ $`=`$ $`{\displaystyle \frac{1}{3}}\theta \sigma _{ij}(\sigma _i^k\sigma _{kj}{\displaystyle \frac{1}{3}}\sigma _{kl}\sigma ^{kl}\delta _{ij})+s_{ij}`$
$`H_{ij}`$ $`=`$ $`3\sigma _{(i}^kn_{j)k}+n_{kl}\sigma ^{kl}\delta _{ij}+{\displaystyle \frac{1}{2}}n_k^k\sigma _{ij},`$
where $`s_{ij}`$ is the same expression that appears in (21.6), see p. 40 of . Observe that in our situation, $`E`$ and $`H`$ are diagonal, since we are interested in the developments obtained in Lemma 21.2. It is natural to normalize $`\stackrel{~}{E}_{ij}=E_{ij}/\theta ^2`$ and similarly for $`H`$. We will denote the diagonal components of $`\stackrel{~}{E}_{ij}`$ by $`\stackrel{~}{E}_i`$. We want to have expressions in $`\mathrm{\Sigma }_+`$, $`\mathrm{\Sigma }_{}`$ and so on, and therefore we compute
$`\stackrel{~}{H}_1`$ $`=`$ $`N_1\mathrm{\Sigma }_++{\displaystyle \frac{1}{\sqrt{3}}}(N_2N_3)\mathrm{\Sigma }_{}`$
$`\stackrel{~}{H}_2`$ $`=`$ $`{\displaystyle \frac{1}{2}}N_2(\mathrm{\Sigma }_++\sqrt{3}\mathrm{\Sigma }_{})+{\displaystyle \frac{1}{2}}(N_3N_1)(\mathrm{\Sigma }_+{\displaystyle \frac{1}{\sqrt{3}}}\mathrm{\Sigma }_{})`$
$`\stackrel{~}{E}_2\stackrel{~}{E}_3`$ $`=`$ $`{\displaystyle \frac{2}{3\sqrt{3}}}\mathrm{\Sigma }_{}(12\mathrm{\Sigma }_+)+(N_2N_3)(N_2+N_3N_1)`$
$`\stackrel{~}{E}_2+\stackrel{~}{E}_3`$ $`=`$ $`{\displaystyle \frac{2}{9}}\mathrm{\Sigma }_+(1+\mathrm{\Sigma }_+){\displaystyle \frac{2}{9}}\mathrm{\Sigma }_{}^2{\displaystyle \frac{2}{3}}N_1^2+{\displaystyle \frac{1}{3}}(N_2N_3)^2+{\displaystyle \frac{1}{3}}N_1(N_2+N_3).`$
Observe that all other components of $`\stackrel{~}{E}_i`$ and $`\stackrel{~}{H}_i`$ can be computed from this, as $`E_{ij}`$ and $`H_{ij}`$ are both traceless.
It is convenient to define the normalized Kretschmann scalar
(22.3)
$$\stackrel{~}{\kappa }=R_{\alpha \beta \gamma \delta }R^{\alpha \beta \gamma \delta }/\theta ^4.$$
The latter object can be expressed as a polynomial in the variables of Wainwright and Hsu. By the above observations and the fact that $`\mathrm{\Omega }=3\mu /\theta ^2`$, we have
$$\stackrel{~}{\kappa }=8[\frac{3}{2}(\stackrel{~}{E}_2+\stackrel{~}{E}_3)^2+\frac{1}{2}(\stackrel{~}{E}_2\stackrel{~}{E}_3)^22\stackrel{~}{H}_1^22\stackrel{~}{H}_2^22\stackrel{~}{H}_1\stackrel{~}{H}_2]+\frac{1}{27}[4+(3\gamma 2)^2]\mathrm{\Omega }^2.$$
We will associate a $`\kappa `$ and a $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$ to a solution to (2.1)-(2.3) in the following way. Since $`\kappa /\theta ^4`$ can be expressed in terms of the variables of Wainwright and Hsu, it is natural to define $`\kappa `$ by this expression multiplied by $`\theta ^4`$, where $`\theta `$ obeys (21.12). There is of course an ambiguity as to the initial value of $`\theta `$, but we are only interested in the asymptotics, and any non-zero value will yield the same conclusion. We associate $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$ to a solution similarly.
###### Lemma 22.1.
The normalized Kretschmann scalar (22.3) is non-zero at the fixed points $`F,P_i^+(II)`$, at the non-special points on the Kasner circle, and at the type I stiff fluid points with $`\mathrm{\Omega }>0`$. Consequently
(22.4)
$$\underset{\tau \mathrm{}}{lim\; sup}|\kappa (\tau )|=\mathrm{}$$
for all solutions to (2.1)-(2.3) which have one such point as an $`\alpha `$-limit point.
Proof. The statement concerning the normalized Kretschmann scalar is a computation. Equation (22.4) is a consequence of this computation, the fact that $`\kappa =\stackrel{~}{\kappa }\theta ^4`$ and the fact that $`\theta \mathrm{}`$ as $`\tau \mathrm{}`$, cf. (21.12). $`\mathrm{}`$
For some non-vacuum Taub type solutions with $`2/3<\gamma <2`$, the following lemma is needed.
###### Lemma 22.2.
Consider a solution to (2.1)-(2.3) with $`\mathrm{\Omega }>0`$ and $`2/3<\gamma <2`$ such that
(22.5)
$$\underset{\tau \mathrm{}}{lim}(\mathrm{\Sigma }_+,\mathrm{\Sigma }_{})=(1,0).$$
Then
$$\underset{\tau \mathrm{}}{lim}\kappa (\tau )=\mathrm{}.$$
Proof. By Proposition 3.1, the solution must satisfy $`\mathrm{\Sigma }_{}=0`$ and $`N_2=N_3`$. Observe that because of (22.5), we have $`\mathrm{\Omega }0`$, since $`\mathrm{\Omega }`$ decays exponentially for $`\mathrm{\Sigma }_+^2`$ large, cf. the proof of Lemma 14.1. Consequently, $`q2`$. One can then prove that for any $`ฯต>0`$, there is a $`T`$ such that
(22.6) $`\mathrm{exp}[(a_\gamma +ฯต)\tau ]\mathrm{\Omega }(\tau )\mathrm{exp}[(a_\gamma ฯต)\tau ]`$
(22.7) $`\mathrm{exp}[(6+ฯต)\tau ]N_1(\tau )\mathrm{exp}[(6ฯต)\tau ]`$
(22.8) $`\mathrm{exp}[(6+ฯต)\tau ][N_1(N_2+N_3)](\tau )\mathrm{exp}[(6ฯต)\tau ]`$
(22.9) $`\mathrm{exp}[(6+ฯต)\tau ]\theta ^2(\tau )\mathrm{exp}[(6ฯต)\tau ]`$
for all $`\tau T`$, where $`a_\gamma =3(2\gamma )`$. However, the constraint can be written
$$(1\mathrm{\Sigma }_+)(1+\mathrm{\Sigma }_+)=\mathrm{\Omega }+\frac{3}{4}N_1^2\frac{3}{2}N_1(N_2+N_3).$$
By (22.6)-(22.8), $`\mathrm{\Omega }`$ will dominate the right hand side, since it is non-zero. Since $`1\mathrm{\Sigma }_+`$ converges to $`2`$, $`1+\mathrm{\Sigma }_+`$ will consequently have to be positive and of the order of magnitude $`\mathrm{\Omega }`$. In particular, for every $`ฯต>0`$ there is a $`T`$ such that
(22.10)
$$\mathrm{exp}[(a_\gamma +ฯต)\tau ](1+\mathrm{\Sigma }_+)(\tau )\mathrm{exp}[(a_\gamma ฯต)\tau ]$$
Observe that since $`a_\gamma <4`$, $`\mathrm{\Omega }\theta ^2`$ and $`(1+\mathrm{\Sigma }_+)\theta ^2`$ both diverge to infinity as $`\tau \mathrm{}`$, by (22.6), (22.9) and (22.10). Other expressions of interest are $`N_1\theta ^2`$ and $`N_1(N_2+N_3)\theta ^2`$. The estimates (22.6)-(22.9) do not yield any conclusions concerning whether they are bounded or not. However, using (21.12), we have
$$N_1(\tau )\theta ^2(\tau )=N_1(0)\theta ^2(0)\mathrm{exp}[_\tau ^0(2+q+4\mathrm{\Sigma }_+)๐s]=$$
$$=N_1(0)\theta ^2(0)\mathrm{exp}[_\tau ^0(2(1+\mathrm{\Sigma }_+)+\frac{1}{2}(3\gamma 2)\mathrm{\Omega }+2\mathrm{\Sigma }_+(1+\mathrm{\Sigma }_+))๐s],$$
which is bounded since all the terms appearing in the integral are integrable by (22.6) and (22.10). A similar argument yields the same conclusion concerning $`N_1(N_2+N_3)\theta ^2`$.
Since the solution is of Taub type, we have $`\stackrel{~}{H}_1=N_1\mathrm{\Sigma }_+`$ and $`\stackrel{~}{H}_2=\stackrel{~}{H}_3=\stackrel{~}{H}_1/2`$. We also have $`\stackrel{~}{E}_2=\stackrel{~}{E}_3`$ and
$$2\stackrel{~}{E}_2=\frac{2}{9}\mathrm{\Sigma }_+(1+\mathrm{\Sigma }_+)\frac{2}{3}N_1^2+\frac{1}{3}N_1(N_2+N_3).$$
Consequently the $`E`$ field blows up and the $`H`$ field remains bounded, and the lemma follows. $`\mathrm{}`$
Finally, we observe that $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$ becomes unbounded in the matter case.
###### Lemma 22.3.
Consider a solution to (2.1)-(2.3) with $`\mathrm{\Omega }>0`$. Then
$$\underset{\tau \mathrm{}}{lim}\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }=\mathrm{}.$$
Remark. How to associate $`\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }`$ to a solution of (2.1)-(2.3) is clarified in the remarks preceding the statement of Lemma 22.1.
Proof. We have
$$\overline{R}_{\alpha \beta }\overline{R}^{\alpha \beta }=\mu ^2+3p^2=[1+3(\gamma 1)^2]\mu ^2=\frac{1}{9}[1+3(\gamma 1)^2]\mathrm{\Omega }^2\theta ^4.$$
But by (2.1) and (21.12), we have
$$\mathrm{\Omega }^2(\tau )\theta ^4(\tau )=\mathrm{\Omega }^2(0)\theta ^4(0)\mathrm{exp}(_\tau ^0(4q+2(3\gamma 2)+4+4q)๐s)=$$
$$=\mathrm{\Omega }^2(0)\theta ^4(0)\mathrm{exp}(3\gamma \tau ),$$
and the lemma follows. $`\mathrm{}`$
###### Lemma 22.4.
Consider a class A development, not of type IX, with $`I=(t_{},t_+)`$ and $`\theta >0`$. Then the corresponding solution to the equations of Wainwright and Hsu has existence interval $``$, and $`tt_\pm `$ corresponds to $`\tau \pm \mathrm{}`$.
Proof. The function $`\theta `$ has to converge to infinity as $`tt_{}`$ for the following reason. Assume it does not. As $`\theta `$ is monotone decreasing, we can assume it to be bounded on $`(t_{},0]`$. By the constraint (21.8), $`\sigma _{ij}`$ and $`\mu `$ are then bounded on $`(t_{},0]`$, so that the same will be true of $`n_{ij}`$ by (21.4) and the fact that $`t_{}>\mathrm{}`$. But then one can extend the solution beyond $`t_{}`$, contradicting the fact that $`I`$ is the maximal existence interval. By (21.7), $`\theta 0`$ as $`t\mathrm{}=t_+`$. Equation (21.10) defines a diffeomorphism $`\stackrel{~}{\tau }:(t_{},t_+)(\tau _{},\tau _+)`$, and we get a solution to the equations of Wainwright and Hsu on $`(\tau _{},\tau _+)`$. By (21.12), we conclude that the statement of the lemma holds. $`\mathrm{}`$
###### Lemma 22.5.
Consider a Bianchi IX class A development with $`I=(t_{},t_+)`$ and $`1\gamma 2`$. According to Lemma 21.6, there is a $`t_0I`$ such that $`\theta >0`$ in $`I_{}=(t_{},t_0)`$ and $`\theta <0`$ in $`I_+=(t_0,t_+)`$. The solution to the equations of Wainwright and Hsu corresponding to the interval $`I_{}`$ has existence interval $`(\mathrm{},\tau _{})`$, and $`tt_{}`$ corresponds to $`\tau \mathrm{}`$. Similarly, $`I_+`$ corresponds to $`(\mathrm{},\tau _+)`$ with $`tt_+`$ corresponding to $`\tau \mathrm{}`$.
Proof. Let us relate the different time coordinates on $`I_{}`$. According to equation (21.10), $`\tau `$ has to satisfy $`dt/d\tau =3/\theta `$. Define $`\stackrel{~}{\tau }(t)=_{t_1}^t\theta (s)/3๐s`$, where $`t_1I_{}`$. Then $`\stackrel{~}{\tau }:I_{}\stackrel{~}{\tau }(I_{})`$ is a diffeomorphism and strictly monotone on $`I_{}`$. Since $`\theta `$ is positive in $`I_{}`$, $`\stackrel{~}{\tau }`$ increases with $`t`$.
Since $`\theta `$ is continuous beyond $`t_0`$, it is clear that $`\stackrel{~}{\tau }(t)\tau _{}`$ as $`tt_0`$. To prove that $`tt_{}`$ corresponds to $`\tau \mathrm{}`$, we make the following observation. One of the expressions $`\theta `$ and $`d\theta /dt`$ is unbounded on $`(t_{},t_1]`$, since if both were bounded the same would be true of $`\sigma _{ij}`$, $`\mu `$ and $`n_{ij}`$ by (21.7) and (21.4) respectively. Then we would be able to extend the solution beyond $`t_{}`$, contradicting the fact that $`I`$ is the maximal existence interval (observe that $`t_{}>\mathrm{}`$ by Lemma 21.8). If $`\stackrel{~}{\tau }`$ were bounded from below on $`I_{}`$, then $`\theta `$ and $`\theta ^{}`$ would be bounded on $`\stackrel{~}{\tau }((t_{},t_1])`$ by Lemma 3.2, and thus $`\theta `$ and $`d\theta /dt`$ would be bounded on $`(t_{},t_1]`$. Thus $`tt_{}`$ corresponds to $`\tau \mathrm{}`$. Similar arguments yield the same conclusion concerning $`I_+`$. $`\mathrm{}`$
## Acknowledgments
This research was supported in part by the National Science Foundation under Grant No. PHY94-07194. Part of this work was carried out while the author was enjoying the hospitality of the Institute for Theoretical Physics, Santa Barbara. The author also wishes to acknowledge the support of Royal Swedish Academy of Sciences. Finally, he would like to express his gratitude to Lars Andersson and Alan Rendall, whose suggestions have improved the article.
|
warning/0006/cond-mat0006272.html
|
ar5iv
|
text
|
# Quantum to classical crossover in the 2D easy-plane XXZ model
## I Introduction
The magnetic properties of low-dimensional quantum spin systems with spin anisotropy, such as the quasi-one-dimensional (1D) cuprates and the quasi-2D high-$`T_c`$ parent compounds , are of growing interest. The $`S=1/2`$ XXZ model
$$=\frac{J}{2}\underset{i,j}{}(S_i^+S_j^{}+\mathrm{\Delta }S_i^zS_j^z)$$
(1)
($`i,j`$ denote nearest-neighbor (NN) sites; throughout we set $`J=1`$) usually serves as the generic model for those systems. Recently, in the ferromagnetic (FM) region ($`1<\mathrm{\Delta }<0`$) of the 1D model a quantum-classical crossover in the longitudinal spin correlators was found by means of exact diagonalization (ED) and a Greenโs-function theory . For the XXZ model on a square lattice, an analytical approach to the spin susceptibility taking into account the magnetic short-range order (SRO) at arbitrary temperatures does not yet exist.
In this contribution the spin correlations in the easy-plane region $`1<\mathrm{\Delta }<1`$ of the 2D XXZ model are examined by both a Greenโs-function theory outlined in the Appendix and by exact finite-cluster diagonalizations of the model (1) on lattices with up to 36 spins using periodic boundary conditions. We mainly focus on the characteristics of a possible quantum to classical crossover in the FM regime. Moreover, for the first time, the complete wave-vector, temperature and $`\mathrm{\Delta }`$ dependences of the static transverse and longitudinal spin susceptibilities are calculated.
## II Ground-state properties
In Fig. 1 our results for the magnetization $`m(\mathrm{\Delta })`$ are compared with available quantum Monte Carlo (QMC) data , where the ED/QMC data for the ground-state energy per site $`\epsilon (\mathrm{\Delta })`$ (inset) is taken as input for the Greenโs-function approach ($`C_{10}^{zz}=\frac{1}{2}\epsilon /\mathrm{\Delta }`$, $`C_{10}^+=\epsilon /2\mathrm{\Delta }C_{10}^{zz}`$).
As can be seen from Fig. 2, the short-ranged correlations calculated analytically are in excellent agreement with our ED data. Let us stress that the finite-size dependence of the ED data is almost negligible by going from a 32- to a 36-site lattice. At $`\mathrm{\Delta }=1`$ the rotational symmetry $`C_๐ซ^+=2C_๐ซ^{zz}`$ is visible. At the quantum critical point $`\mathrm{\Delta }=1`$ we have $`C_{๐ซ,\stackrel{~}{}}^+=2C_๐ซ^{zz}=1/6`$ (cf. Eq. (18)). The non-analytical limiting behavior $`lim_{\mathrm{\Delta }1^+}C_๐ซ^{zz}=0`$ results from both the QMC and ED data (obtained in the subspace with total spin projection $`S^z=0`$).
The static spin susceptibilities $`\chi _๐ช^\nu (\mathrm{\Delta })`$ are depicted in Fig. 3. In the FM region, for sufficiently low $`\mathrm{\Delta }`$ values, $`\chi _๐ช^{zz}`$ shows a maximum at $`๐ช=0`$ being a precursor of the FM instability (in the $`zz`$-correlators) at $`\mathrm{\Delta }=1`$. Note that $`(\chi _๐^+)^1=0`$, reflecting the transverse long-range order (LRO) at $`T=0`$, by Eq. (17) corresponds to $`(\chi _{0,\stackrel{~}{}}^+)^1=0`$. In the antiferromagnetic (AFM) region $`0<\mathrm{\Delta }<1`$ the maximum in $`\chi _๐ช^{zz}`$ at $`๐ช=๐`$ is indicative of the longitudinal AFM LRO at $`\mathrm{\Delta }1`$.
Finally, in Fig. 4 we show the longitudinal spin-wave spectrum $`\omega _๐ช^{zz}`$ (cf. Eq. (11)). For $`q|๐ช|1`$ we have $`\omega _๐ช^{zz}=c_s^{zz}q`$, where the spin-wave velocity $`c_s^{zz}`$ increases with $`\mathrm{\Delta }`$ over the whole easy-plane region. The minimum in $`\omega _๐ช^{zz}`$ at $`๐ช=๐`$ in the AFM region corresponds to the maximum in $`\chi _๐ช^{zz}`$ (cf. Fig. 3 a) and reflects the increase of the longitudinal AFM SRO with $`\mathrm{\Delta }`$ (see also $`C_๐ซ^{zz}(\mathrm{\Delta })`$ in Fig. 2).
## III Finite-temperature results
The temperature dependence of the short-ranged longitudinal spin correlations is displayed in Fig. 5. Again the analytical results agree remarkably well with the ED data. In the FM region, for the first time in the 2D model, we observe the so-called โsign-changingโ effect which was found numerically in the 1D model and later on reproduced by our Greenโs-function calculations . That is, at fixed separation $`r`$ and with increasing temperature or at fixed temperature and with increasing $`r`$, $`C_๐ซ^{zz}`$ changes sign from negative to positive values. The temperature $`T_0(\mathrm{\Delta },๐ซ)`$ where $`C_๐ซ^{zz}(T_0(\mathrm{\Delta },๐ซ),\mathrm{\Delta })=0`$ are given in Table I. As in the 1D case, $`T_0`$ at fixed $`\mathrm{\Delta }`$ decreases with increasing $`r`$. However, compared to the 1D case , our analytical results are in much better agreement with the ED data. The sign change of $`C_๐ซ^{zz}`$ may be interpreted as a quantum to classical crossover because with increasing temperature the system behaves more classically, i.e., it becomes dominated by the potential energy (negative $`\mathrm{\Delta }`$ term of the Hamiltonian favoring the parallel alignment of two spins). In the AFM region we obtain the expected alternating signs of $`C_๐ซ^{zz}`$ corresponding to the longitudinal AFM SRO.
In Fig. 6 various susceptibilities $`\chi _๐ช^\nu `$ at $`๐ช=0,๐`$ are plotted as functions of $`T`$ and compared with numerical data. For $`\mathrm{\Delta }=0.5`$ the longitudinal and transverse uniform susceptibilities are in reasonable agreement with the QMC results and our ED data (the up- and downturn at lower temperatures is a finite-size effect). The increase of $`\chi _0^\nu (T)`$, the maximum near the exchange energy ($`J=1`$), and the crossover to the Curie-Weiss law are due to the decrease of AFM SRO with increasing temperature. On the other hand, the staggered susceptibility $`\chi _๐^{zz}`$ is enhanced as compared with $`\chi _0^{zz}`$ by the longitudinal AFM SRO. In the FM region (Fig. 6 b, $`\mathrm{\Delta }=0.5`$) the maximum in $`\chi _0^{zz}`$, where the analytical and numerical results yield nearly the same position, may be explained as a combined SRO and sign changing effect as discussed for the 1D model in Ref. . Contrary to the AFM region, $`\chi _๐^{zz}`$ is suppressed as compared with $`\chi _0^{zz}`$ which is caused by the FM correlations above $`T_0`$.
The temperature dependence of $`\chi _0^+=\chi _{๐,\stackrel{~}{}}^+`$ may be explained again as a SRO effect.
Here, the transverse FM SRO results in a spin stiffness against the orientation of the transverse spin components along a staggered field perpendicular to the $`z`$-direction, so that $`\chi _{๐,\stackrel{~}{}}^+`$ is suppressed at low temperatures and exhibits a maximum.
## IV Summary
To resume, we presented a Greenโs-function theory of magnetic LRO and SRO in the 2D easy-plane XXZ model which allows the complete calculation of all static magnetic properties in excellent agreement with numerical diagonalization data. In particular, in the FM region we found a quantum to classical crossover in the longitudinal spin correlations. We conclude that our approach is promising to be applied to other anisotropic spin models, such as the quasi-2D XXZ model for the parent compounds of high-$`T_c`$ superconductors.
## Appendix: Greenโs-function theory
The spin susceptibilities $`\chi _๐ช^+(\omega )=S_๐ช^+;S_๐ช^{}_\omega `$ and $`\chi _๐ช^{zz}(\omega )=S_๐ช^z;S_๐ช^z_\omega `$, expressed in terms of two-time retarded commutator Greenโs functions, are determined by the projection method, developed, for the XXZ chain, in Ref. . Taking the two-operator basis $`(S_๐ช^+,i\dot{S}_๐ช^+)^T`$ and $`(S_๐ช^z,i\dot{S}_๐ช^z)^T`$ we obtain
$$\chi _๐ช^\nu (\omega )=\frac{M_๐ช^\nu }{\omega ^2(\omega _๐ช^\nu )^2};\nu =+,zz,$$
(2)
with
$`M_๐ช^+`$ $`=`$ $`4[C_{10}^+(1\mathrm{\Delta }\gamma _๐ช)+2C_{10}^{zz}(\mathrm{\Delta }\gamma _๐ช)],`$ (3)
$`M_๐ช^{zz}`$ $`=`$ $`4C_{10}^+(1\gamma _๐ช),`$ (4)
$`C_{nm}^\nu C_๐ซ^\nu `$, $`C_๐ซ^+=S_0^+S_๐ซ^{}`$, $`C_๐ซ^{zz}=S_0^zS_๐ซ^z`$, $`๐ซ=n๐_x+m๐_y`$, and $`\gamma _๐ช=(\mathrm{cos}q_x+\mathrm{cos}q_y)/2`$. The spin correlators are obtained from Eq. (2) as
$$C_๐ซ^\nu =\frac{1}{N}\underset{๐ช}{}\frac{M_๐ช^\nu }{2\omega _๐ช^\nu }[1+2p(\omega _๐ช^\nu )]\text{e}^{i\mathrm{๐ช๐ซ}},$$
(5)
where $`p(\omega _๐ช^\nu )=(\text{e}^{\omega _๐ช^\nu /T}1)^1`$. The spectra $`\omega _๐ช^\nu `$, calculated in the approximations $`\ddot{S}_๐ช^+=(\omega _๐ช^+)^2S_๐ช^+`$ and $`\ddot{S}_๐ช^z=(\omega _๐ช^{zz})^2S_๐ช^z`$ introducing vertex parameters $`\alpha _i^\nu `$ ($`i=1,2`$), are given by
$`(\omega _๐ช^+)^2`$ $`=`$ $`[(1+2\alpha _2^+(C_{20}^++2C_{11}^+)](1\mathrm{\Delta }\gamma _๐ช)`$ (9)
$`+\mathrm{\Delta }(1+4\alpha _2^+(C_{20}^{zz}+2C_{11}^{zz})](\mathrm{\Delta }\gamma _๐ช)`$
$`+2\alpha _1^+[C_{10}^+(4\mathrm{\Delta }\gamma _๐ช^2\mathrm{\Delta }3\gamma _๐ช)`$
$`+2C_{10}^{zz}(4\gamma _๐ช^213\mathrm{\Delta }\gamma _๐ช)],`$
$`(\omega _๐ช^{zz})^2`$ $`=`$ $`2(1\gamma _๐ช)[1+2\alpha _2^{zz}(C_{20}^++2C_{11}^+)`$ (11)
$`2\mathrm{\Delta }\alpha _1^{zz}C_{10}^+(1+4\gamma _๐ช)].`$
In the easy-plane region $`1<\mathrm{\Delta }<1`$, the long-range order at $`T=0`$ is reflected in our theory by $`\omega _๐^+=0`$ \[$`๐=(\pi ,\pi )`$\]. Accordingly, the condensation part $`C\text{e}^{i\mathrm{๐๐ซ}}`$ is separated from $`C_๐ซ^+`$ (cf. Eq. (5), and the magnetization $`m`$ is calculated as
$$m^2=\frac{1}{N}\underset{๐ซ}{}C_๐ซ^+\text{e}^{i\mathrm{๐๐ซ}}=C.$$
(12)
The parameters $`\alpha _1^\nu (T)`$ are determined from the sum rules $`C_{00}^+=1/2`$ and $`C_{00}^{zz}=1/4`$. To obtain $`\alpha _2^\nu (T)`$ we adjust $`C_{10}^\nu (T=0)`$ taken from our ED data and assume, as additional conditions for the calculation of $`\chi _๐ช^{zz}(\omega )`$ and $`\chi _๐ช^+(\omega )`$, temperature independent ratios
$$R^{zz}=\frac{\alpha _2^{zz}(T)1}{\alpha _1^{zz}(T)1}$$
(13)
and
$`R_>^+`$ $`=`$ $`{\displaystyle \frac{\alpha _2^+(T)1}{\alpha _1^+(T)1}}\text{for}\mathrm{\Delta }>0,`$ (14)
$`R_<^+`$ $`=`$ $`{\displaystyle \frac{\alpha _2^+(T)1}{\alpha _1^{zz}(T)1}}\text{for}\mathrm{\Delta }<0,`$ (15)
respectively. For the discussion it is useful to perform the unitary transformation which rotates the spins on the sublattice B around the $`z`$-axis by the angle $`\pi `$, $`\stackrel{~}{๐}_i=๐ฐ^+๐_i๐ฐ`$ with $`๐ฐ=_{lB}2S_l^z`$. We get $`\stackrel{~}{S}_i^{x,y}=\text{e}^{i\mathrm{๐๐ซ}_i}S_i^{x,y}`$, $`\stackrel{~}{S}_i^z=S_i^z`$ and
$$\stackrel{~}{}=\frac{1}{2}\underset{i,j}{}(S_i^+S_j^{}+\mathrm{\Delta }S_i^zS_j^z).$$
(16)
Due to $`A_{}=\stackrel{~}{A}_\stackrel{~}{}`$ for any operator $`A`$, we obtain the relations
$`\chi _{๐ช,}^+(\omega )`$ $`=`$ $`\chi _{๐ค,\stackrel{~}{}}^+(\omega );๐ค=๐ช๐,`$ (17)
$`C_{๐ซ,}^+`$ $`=`$ $`\text{e}^{i\mathrm{๐๐ซ}}C_{๐ซ,\stackrel{~}{}}^+,`$ (18)
$`\chi _{๐ช,}^{zz}(\omega )=\chi _{๐ช,\stackrel{~}{}}^{zz}(\omega )`$, and $`C_{๐ซ,}^{zz}=C_{๐ซ,\stackrel{~}{}}^{zz}`$. As shown in Ref. , the rotational symmetry at $`\mathrm{\Delta }=\pm 1`$ is preserved by our theory.
|
warning/0006/cond-mat0006263.html
|
ar5iv
|
text
|
# A toy model for molecular condensates in Bose gases
## Abstract
The occurrence of a molecular Bose-Einstein condensate is studied for an atomic system near a zero energy resonance of the binary scattering process, with a large and positive scattering length. The interaction potential is modeled by a pseudo-potential having one bound state. Using a variational Gaussian ansatz for the $`N`$-body density operator, we discuss the thermodynamic properties at low temperature and the relative stability of the system towards the formation of an atomic Bose-Einstein condensate. We also derive an approximate Gross-Pitaevskii equation for the molecular condensate leading to the prediction of a Bogoliubov spectrum.
The atomic Bose-Einstein condensation has been extensively studied during past years both theoretically and experimentally . In this Letter, we consider the different case of a molecular Bose-Einstein condensate. The realization of such a situation is a challenge in the field of low temperature physics. Already cold gases of molecules have been produced . No evidence for a molecular condensate has been obtained yet, although the stimulated Raman technique used in , forming molecules directly from an atomic condensate, is promising.
Another scenario is possible for the formation of the molecular condensate. Consider the case of a so-called zero-energy resonance in the two-body scattering process, where the positive scattering length $`a>0`$ is much larger than the effective range $`r_e`$ of the interaction potential . In this case the two-body potential supports a $`s`$-wave bound state $`\varphi _0`$ with a spatial extension $`a`$ much larger than the other $`s`$-wave bound states and with a much smaller binding energy $`E_0=\mathrm{}^2/(ma^2)`$. One then would rely simply on three-body collisions between atoms of a trapped atomic condensate to produce molecules in the state $`\varphi _0`$. As shown in the rate of formation of molecules scales as $`a^4`$ in the limit of large $`a`$ and mainly leads to formation of molecules in the highest $`s`$-wave two-body channel that is in the state $`\varphi _0`$. From energy conservation one finds that the molecules produced in $`\varphi _0`$ have a kinetic energy $`E_0/3`$ when the initial atomic wavevectors $`\stackrel{}{k}`$ are such that $`ka1`$. The requirement that the molecules formed in state $`\varphi _0`$ remain in the trap imposes a depth of the trapping potential larger than $`E_0`$. For a modest trap depth of 10 $`\mu `$K achievable in an optical trap and for the mass of Rubidium one finds that $`a`$ has to be larger than 445 Bohr radii. Such high values of $`a`$ may be obtained using a Feshbach resonance induced by a magnetic field .
In this Letter we assume that a condensate of molecules in the diatomic bound state $`\varphi _0`$ has been formed. We describe the molecular condensate at the atomic level: we take as a starting point a model Hamiltonian for interacting atoms and we use a Gaussian variational ansatz for the $`N`$body density operator including only the effect of binary interactions between atoms. In the low density regime ($`na^31`$ where $`n`$ is the density of atoms) we determine (i) the properties of the molecular condensate at thermal equilibrium and (ii) the response of the molecular condensate to a small perturbation of the trapping potential. Our results for both cases correspond formally to the known properties of a weakly interacting Bose-Einstein condensate of particles of mass $`2m`$ having a coupling constant $`6g`$, where $`m`$ is the atomic mass and $`g=4\pi \mathrm{}^2a/m`$ is the atomic coupling constant. At higher density ($`na^3>\pi /192`$) we find that an atomic condensate forms. To model the binary atomic interaction potential we use the pseudo-potential $`V`$ defined by the following action on a two-body atomic wavefunction $`\psi `$:
$$\stackrel{}{r}_1,\stackrel{}{r}_2|V|\psi =g\delta (\stackrel{}{r})_r\left[r\psi (\stackrel{}{R}\stackrel{}{r}/2,\stackrel{}{R}+\stackrel{}{r}/2)\right],$$
(1)
where we have introduced the coordinates of the center of mass and of the relative motion:
$$\stackrel{}{R}=\frac{\stackrel{}{r}_1+\stackrel{}{r}_2}{2},\stackrel{}{r}=\stackrel{}{r}_1\stackrel{}{r}_2\text{and}r=\stackrel{}{r}.$$
(2)
This model potential discussed in has been used to extend the BCS theory to inhomogeneous atomic systems . As we now see, it has the ability to capture the essential feature of a general binary scattering problem near a zero-energy resonance. Consider indeed the relative motion of two atoms described by the hamiltonian
$$_r=\frac{\mathrm{}^2}{m}\mathrm{\Delta }_\stackrel{}{r}+V.$$
(3)
Contrary to the case of the usual contact interaction, the scattering of two atoms interacting with the potential (1) is a well defined problem. The scattering amplitude for a relative wavevector $`\stackrel{}{k}`$ is given by $`a/(1+ika)`$ reproducing the known universal Lorentzian shape of the scattering cross section near a zero energy resonance . In addition to the usual scattering states, the pseudo-potential for $`a>0`$ leads to the asymptotic form of the $`s`$-wave bound state of energy $`E_0`$:
$$\varphi _0(\stackrel{}{r})=\frac{1}{r\left(2\pi a\right)^{1/2}}\mathrm{exp}\left(\frac{r}{a}\right).$$
(4)
This property of the pseudo-potential reproduces the universal behaviour for a general interaction potential close to a zero-energy resonance, already mentioned in the introduction. It plays an essential role in our approach as it allows the formation of pairs of atoms, that is binary molecules, while keeping the mathematical simplicity of a zero range model potential ($`r_e=0`$).
The goal of this paper is to study the properties of a molecular condensate, in a symmetry breaking approach, this corresponds to $`\widehat{\mathrm{\Psi }}=0`$ and $`\widehat{\mathrm{\Psi }}\widehat{\mathrm{\Psi }}0`$, where $`\widehat{\mathrm{\Psi }}`$ is the atomic field operator. However it is dangerous to exclude a priori the coexistence of an atomic and a molecular condensate. We therefore use the more general symmetry breaking prescription by splitting the field operator in a classical part and a quantum fluctuation part: $`\widehat{\mathrm{\Psi }}=\mathrm{\Phi }+\widehat{\varphi }`$, with $`\widehat{\varphi }=0`$. We choose a Gaussian ansatz for the many-body density matrix $`D=Z^1\mathrm{exp}\beta K`$ with
$`K`$ $`=`$ $`{\displaystyle d^3r_1d^3r_2\widehat{\varphi }^{}(\stackrel{}{r}_1)h(\stackrel{}{r}_1,\stackrel{}{r}_2)\widehat{\varphi }(\stackrel{}{r}_2)}`$ (6)
$`+{\displaystyle \frac{1}{2}}{\displaystyle }d^3r_1d^3r_2[\widehat{\varphi }(\stackrel{}{r}_1)\mathrm{\Delta }(\stackrel{}{r}_1,\stackrel{}{r}_2)\widehat{\varphi }(\stackrel{}{r}_2)+\mathrm{h}.\mathrm{c}.].`$
There are three variational fields in this theory: $`h,\mathrm{\Delta }`$ and $`\mathrm{\Phi }`$. At thermal equilibrium, the grand potential $`E\mu NTS`$ is minimum so that the functions $`h(\stackrel{}{r}_1,\stackrel{}{r}_2)`$ and $`\mathrm{\Delta }(\stackrel{}{r}_1,\stackrel{}{r}_2)`$ may be expressed in terms of the one-body density matrix $`\overline{\rho }(\stackrel{}{r}_1,\stackrel{}{r}_2)=\widehat{\varphi }^{}(\stackrel{}{r}_2)\widehat{\varphi }(\stackrel{}{r}_1)`$ and of the pairing function $`\overline{\kappa }(\stackrel{}{r}_1,\stackrel{}{r}_2)=\widehat{\varphi }(\stackrel{}{r}_1)\widehat{\varphi }(\stackrel{}{r}_2)`$, while $`\mathrm{\Phi }(\stackrel{}{r})`$ verifies a partial differential equation
$`h(\stackrel{}{r}_1,\stackrel{}{r}_2)`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}(\stackrel{}{}^2\delta )(\stackrel{}{r})+(2gn(\stackrel{}{R})\mu )\delta (\stackrel{}{r})`$ (7)
$`\mathrm{\Delta }(\stackrel{}{r}_1,\stackrel{}{r}_2)`$ $`=`$ $`g\delta (\stackrel{}{r})[\overline{\kappa }_{\mathrm{reg}}(\stackrel{}{R})+\mathrm{\Phi }^2(\stackrel{}{R})]`$ (8)
$`{\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\mathrm{\Phi }`$ $`+`$ $`[g(2n|\mathrm{\Phi }|^2)\mu ]\mathrm{\Phi }+g\overline{\kappa }_{\mathrm{reg}}\mathrm{\Phi }^{}=0`$ (9)
In these equations, we have introduced the atomic density at a point $`\stackrel{}{R}`$
$$n(\stackrel{}{R})=|\mathrm{\Phi }|^2(\stackrel{}{R})+\overline{\rho }(\stackrel{}{R},\stackrel{}{R}),$$
(10)
and the regular part of the pairing function
$$\overline{\kappa }_{\mathrm{reg}}(\stackrel{}{R})=\underset{r0}{lim}_r\left[r\overline{\kappa }(\stackrel{}{R}\stackrel{}{r}/2,\stackrel{}{R}+\stackrel{}{r}/2)\right].$$
(11)
We assume here that the atoms are in a cubic box of size $`L`$ with periodic boundary conditions , so that $`\mathrm{\Phi },n,\kappa _{\mathrm{reg}}`$ do not depend on position. We expand the field operator on plane waves using a Bogoliubov transform:
$$\widehat{\varphi }(\stackrel{}{r})=\frac{1}{L^{3/2}}\underset{\stackrel{}{k}}{}\widehat{b}_\stackrel{}{k}u_k\mathrm{exp}(i\stackrel{}{k}\stackrel{}{r})+\widehat{b}_\stackrel{}{k}^{}v_k^{}\mathrm{exp}(i\stackrel{}{k}\stackrel{}{r}).$$
(12)
The commutation relations of the bosonic annihilation operators $`\widehat{b}_\stackrel{}{k}`$โs lead to the normalization of the modes amplitudes $`|u_k|^2|v_k|^2=1`$. We search the $`(u_k,v_k)`$โs so that $`K`$ is a sum of decoupled harmonic oscillators:
$$K=K_0+\underset{\stackrel{}{k}}{}\mathrm{}\omega _k\widehat{b}_\stackrel{}{k}^{}\widehat{b}_\stackrel{}{k}.$$
(13)
From the equilibrium conditions (7,8), we find that each $`(u_k,v_k)`$ is the eigenvector of the system:
$`h_ku_k+g(\mathrm{\Phi }^2+\overline{\kappa }_{\mathrm{reg}})v_k`$ $`=`$ $`\mathrm{}\omega _ku_k`$ (14)
$`h_kv_k+g(\mathrm{\Phi }^2+\overline{\kappa }_{\mathrm{reg}})u_k`$ $`=`$ $`\mathrm{}\omega _kv_k`$ (15)
with the notation $`h_k={\displaystyle \frac{\mathrm{}^2k^2}{2m}}+2gn\mu `$. A simple algebra gives the expressions for the eigenvectors and the spectrum $`\{\mathrm{}\omega _k\}`$ :
$`v_k^2`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[{\displaystyle \frac{h_k}{\mathrm{}\omega _k}}1\right]`$ (16)
$`\mathrm{}\omega _k`$ $`=`$ $`\left[h_k^2g^2(\mathrm{\Phi }^2+\overline{\kappa }_{\mathrm{reg}})^2\right]^{1/2}`$ (17)
The pairing function is given by
$$\overline{\kappa }(\stackrel{}{r}_1,\stackrel{}{r}_2)=\frac{g(\overline{\kappa }_{\mathrm{reg}}+\mathrm{\Phi }^2)}{2L^3}\underset{\stackrel{}{k}}{}\frac{1+2f_k}{\mathrm{}\omega _k}\mathrm{exp}(i\stackrel{}{k}\stackrel{}{r}),$$
(18)
and the one-body density matrix
$$\overline{\rho }(\stackrel{}{r}_1,\stackrel{}{r}_2)=\frac{1}{L^3}\underset{\stackrel{}{k}}{}\left[\left(2f_k+1\right)v_k^2+f_k\right]\mathrm{exp}(i\stackrel{}{k}\stackrel{}{r}).$$
(19)
In Eqs.(18,19), $`f_k`$ is the Bose occupation factor $`f_k=[\mathrm{exp}(\beta \mathrm{}\omega _k)1]^1`$.
All the equilibrium properties may be expressed in terms of $`(n,T)`$. For this purpose, we have to determine the three unknown parameters $`(\mu ,\mathrm{\Phi },\overline{\kappa }_{\mathrm{reg}})`$. In the presence of an atomic condensate, $`\mathrm{\Phi }`$ does not vanish and a first relation is given by Eq.(9)
$$\mu =g[2n+\overline{\kappa }_{\mathrm{reg}}\mathrm{\Phi }^2].$$
(20)
A second one is obtained from Eq.(10), by setting $`\stackrel{}{r}=0`$ in Eq.(19)
$$n=\mathrm{\Phi }^2+\frac{1}{L^3}\underset{\stackrel{}{k}}{}\left[\left(2f_k+1\right)v_k^2+f_k\right].$$
(21)
The third equation is obtained by extracting the regular part of Eq.(18) as in ; this leads to
$$\overline{\kappa }_{\mathrm{reg}}=\frac{g(\mathrm{\Phi }^2+\overline{\kappa }_{\mathrm{reg}})}{L^3}\underset{\stackrel{}{k}}{}\left[\frac{m}{\mathrm{}^2k^2}\frac{1+2f_k}{2\mathrm{}\omega _k}\right].$$
(22)
In what follows, we consider the thermodynamical limit so that the sums over $`\stackrel{}{k}`$ are replaced by integrals, we also restrict to the case of a vanishing temperature $`T=0`$, so that the Bose occupation factors $`f_k`$ is zero.
First, we note that the stability of the ground state imposes real values for the spectrum $`\{\mathrm{}\omega _k\}`$, this implies $`\overline{\kappa }_{\mathrm{reg}}<0`$ and from Eq.(22), we find $`|\overline{\kappa }_{\mathrm{reg}}|>|\mathrm{\Phi }|^2`$. In particular for a vanishing number of atoms in the condensate $`\mathrm{\Phi }=0`$, we get from Eqs.(20,21,22):
$$\overline{\kappa }_{\mathrm{reg}}^\mathrm{c}=\frac{\pi }{64a^3},n^\mathrm{c}=\frac{\pi }{192a^3}.$$
(23)
This value of the density determines the threshold of coexistence of an atomic condensate ($`\mathrm{\Phi }0`$) with a molecular one. For $`n>n^c`$, the model predicts the coexistence, a result already obtained in a slightly different approach in . In this high density regime, our modelization is questionable and we do not pursue the study in this range anymore.
For $`n<n^\mathrm{c}`$, there is no atomic condensate: $`\mathrm{\Phi }`$ is identically zero so that Eq.(20) does not hold in this regime and equilibrium properties are deduced from Eqs.(21,22) only. We now wish to check that the low density regime corresponds indeed to a molecular condensate. For this purpose, we suppose that the chemical potential tends to a finite negative value, hence $`|\mu |gn,g|\overline{\kappa }_{\mathrm{reg}}|`$ and $`\mathrm{}\omega _k\mathrm{}^2k^2/(2m)\mu `$. From Eq.(22) we find the lowest order approximation
$$\mu \frac{E_0}{2},$$
(24)
which is finite indeed, and negative. This result is enlightening: $`E_0`$ is just the binding energy of a pair of atoms in the bound state Eq.(4) so that the value Eq.(24) of the chemical potential corresponds to that of an ideal gas of molecules. Since we are at zero temperature this gas of pairs is actually a molecular Bose-Einstein condensate.
This interpretation is confirmed by the lowest order approximation to the pairing function $`\kappa `$ in the low density limit. One first calculates $`\overline{\kappa }_{\mathrm{reg}}`$ from Eq.(21) using the lowest order approximation $`v_kg\overline{\kappa }_{\mathrm{reg}}(E_0\mathrm{}^2k^2/m)^1`$ :
$$\overline{\kappa }_{\mathrm{reg}}\frac{1}{a^3}\left(\frac{na^3}{2\pi }\right)^{1/2},$$
(25)
then we calculate the integral over $`\stackrel{}{k}`$ in Eq.(18) and in Eq.(19) to obtain the lowest order contributions to the pairing function and to the one-body density matrix:
$$\overline{\kappa }(\stackrel{}{r}_1,\stackrel{}{r}_2)\sqrt{n}\varphi _0(r),\rho (\stackrel{}{r}_1,\stackrel{}{r}_2)n\mathrm{exp}(\frac{r}{a}).$$
(26)
This expression of $`\overline{\kappa }(\stackrel{}{r}_1,\stackrel{}{r}_2)`$ clearly shows that the pairing function describe the spatial structure of two atoms linked in the molecular bound state $`\varphi _0`$. Similar results have been obtained in for a different model potential.
For a small but finite gaseous parameter $`na^3`$ the molecular condensate is not an ideal gas but rather a weakly interacting Bose gas, with an effective coupling constant $`g_{\mathrm{mol}}`$ between the molecules. We derive this coupling constant by calculating the first correction to the expression Eq.(24) for the chemical potential. We expand the expression Eq.(17) to first order in $`na^3`$ and substitute the result in Eq.(22). This leads to
$$\mu =\frac{E_0}{2}(112\pi na^3+\mathrm{}).$$
(27)
On the other hand the molecular chemical potential $`\mu _{\mathrm{mol}}`$, equal to twice the atomic chemical potential $`\mu `$, is given in the usual mean-field approach for condensates by
$$\mu _{\mathrm{mol}}=E_0+g_{\mathrm{mol}}n_{\mathrm{mol}},$$
(28)
where $`n_{\mathrm{mol}}=n/2`$ is the density of molecules. We therefore deduce for the coupling constant between molecules :
$$g_{\mathrm{mol}}=6g.$$
(29)
As $`g_{\mathrm{mol}}`$ is positive the molecular condensate is stable with respect to a spatial collapse.
The value of the molecular coupling constant $`g_{\mathrm{mol}}`$ can also be obtained from the response of the gas to a time dependent perturbation. It is important to check that the corresponding value coincides with the static prediction Eq.(29). Imagine that one perturbs the system from thermodynamical equilibrium by applying an external potential on the atoms for a finite time interval $`[0,\tau ]`$. We describe the evolution of the gas by a time dependent Gaussian ansatz for the many-body density operator . In that case, as in , we deduce from the Heisenberg equation for the field operator and from Wickโs theorem the time evolution of the pairing function, written here for convenience for $`t>\tau `$:
$`[i\mathrm{}_t`$ $``$ $`{\displaystyle \frac{\mathrm{}^2}{4m}}\mathrm{\Delta }_\stackrel{}{R}+2(gn(\stackrel{}{r}_1)+gn(\stackrel{}{r}_2)\mu )+_r]\overline{\kappa }(\stackrel{}{r}_1,\stackrel{}{r}_2)`$ (30)
$`=`$ $`g\overline{\kappa }_{\mathrm{reg}}(\stackrel{}{r}_1)\rho ^{}(\stackrel{}{r}_1,\stackrel{}{r}_2)g\overline{\kappa }_{\mathrm{reg}}(\stackrel{}{r}_2)\rho (\stackrel{}{r}_1,\stackrel{}{r}_2).`$ (31)
We assume that the applied perturbation varies very slowly spatially at the scale of the scattering length $`a`$. Thus the pairing function has negligible components on the scattering states and can be assumed to have the same $`r`$ dependence as $`\varphi _0`$. We therefore take the Local Density Approximation:
$$\overline{\kappa }(\stackrel{}{r}_1,\stackrel{}{r}_2)=\left[n(\stackrel{}{R},t)\right]^{1/2}\varphi _0(r)\mathrm{exp}\left[iS(\stackrel{}{R},t)\right].$$
(32)
We close equation (31) with the Local Density Approximation for the one-body density matrix:
$$\rho (\stackrel{}{r}_1,\stackrel{}{r}_2)=n(\stackrel{}{R},t)\mathrm{exp}\left(\frac{r}{a}\right).$$
(33)
A simple projection of Eq.(31) on the bound state (4) leads to the Gross-Pitaevskii equation
$$i\mathrm{}_t\psi _P=\left(\frac{\mathrm{}^2}{4m}\mathrm{\Delta }_\stackrel{}{R}+6g|\psi _P|^22\mu +E_0\right)\psi _P,$$
(34)
where we have introduced the macroscopic wave function $`\psi _P(\stackrel{}{R},t)=\left(n/2\right)^{1/2}\mathrm{exp}(iS)`$ describing the molecular condensate. We note that this equation, confirms in a direct way our previous finding on $`\mu `$ at equilibrium (Eq.(27)). From the linear analysis of Eq.(34), we predict a Bogoliubov spectrum for the molecular condensate different from the atomic condensate:
$$\mathrm{}\omega _P(k)=\left(\frac{\mathrm{}^2k^2}{4m}\right)^{1/2}\left(\frac{\mathrm{}^2k^2}{4m}+6gn\right)^{1/2}.$$
(35)
Measurement of this spectrum could be used as an experimental evidence for the condensation of pairs.
As a conclusion let us stress three points. First, one word about the temperature. Indeed, the critical temperature at which the condensate of molecules (mass $`2m`$, density $`n/2`$) forms is
$$k_BT_c=\frac{\pi \mathrm{}^2}{m}\left[\frac{n}{2\zeta (3/2)}\right]^{2/3}.$$
(36)
In the low density regime $`k_BT_c|E_0|`$, so that condensation of pairs can occur without any thermal dissociation. Second, it would be interesting to test the prediction on the coupling constant ($`6g`$) by a direct calculation of the scattering of two molecules of the type considered here. Finally, our model does not describe the relative stability of this molecular condensate toward the formation of molecules in deeper bound levels. An evaluation of the creation rate of deep bound states by collision of an atom with one molecule would be a relevant complement to this analysis.
The author thanks Yvan Castin for helpful contribution to this work. The author also wishes to thank D. Vautherin for stimulating discussions and the members of the LPTL for their warm welcome in the laboratory.
|
warning/0006/astro-ph0006264.html
|
ar5iv
|
text
|
# Detailed Analysis of Early to Late-Time Spectra of Supernova 1993J
## 1 Introduction
Supernova 1993J presented a great opportunity in the study of supernovae (SNe). It occurred in the nearby galaxy M81 (NGC 3031; $`d`$ = 3.6 Mpc; Freedman et al. 1994) and reached a maximum brightness of $`m_V=10.8`$ mag (e.g., Richmond et al. 1996). This allowed extremely detailed observations to be taken over a long time interval. The fact that SN 1993J underwent a transformation from a Type II to a Type IIb, strengthening the link between SNe II and SNe Ib as core-collapse events, only enhanced the importance of this object (see, e.g., Wheeler & Filippenko 1996; Matheson et al. 2000, hereinafter Paper I, for detailed discussions of the nature of SN 1993J).
We have previously presented analyses of the spectra of SN 1993J obtained during its first year (Filippenko, Matheson, & Ho 1993, hereinafter FMH93; Filippenko, Matheson, & Barth 1994, hereinafter FMB94), while Paper I shows our complete low-resolution spectroscopic data set. The early spectra and the transition from SN II to SN IIb are covered by FMH93, while the onset of the nebular phase and the beginning of signs of circumstellar interaction are discussed by FMB94. In this paper, we take a closer look at the details of some of the earlier spectra and we present the analysis of late-time spectra of SN 1993J. The clumpy nature of SN 1993J, as revealed in the details of spectra from early to nebular times, is presented in ยง2. The late-time spectra and their evidence of circumstellar interaction are discussed in ยง3; we summarize our conclusions in ยง4. In the Appendix, we discuss the potential impact of telluric absorption on the interpretation of the spectra of bright SNe and the implications this has for future SN studies.
All of the low-dispersion spectra discussed herein were obtained at the Lick and Keck Observatories. In addition, a single high-dispersion spectrum was obtained using the Keck I 10-m telescope. The details of the observations, including instrumental configuration, dates of observation, and observational parameters, are presented in Paper I. We follow Lewis et al. (1994) in adopting 1993 March 27.5 UT (JD 2,449,074) as the date of explosion.
## 2 Clumps
### 2.1 Clumps in Other Supernovae
The mottled appearance of supernova remnants (SNRs) has often been interpreted as evidence for clumps in the ejecta of SNe. For example, fast-moving features have been identified in Cas A for quite some time (e.g., Baade & Minkowski 1954), and modern studies continue to interpret the structure of Cas A as the result of clumpy SN ejecta (e.g., Anderson et al. 1994). The discovery of metal-enriched fragments outside the boundary of the Vela SNR (Aschenbach et al. 1995), a high-velocity โbulletโ in Vela (Strom et al. 1995), and fast-moving oxygen filaments in Puppis A (and other remnants; Winkler & Kirshner 1985, and references therein) all support the contention that SN ejecta are clumped. There is also a theoretical underpinning for the expectation of clumps in the ejecta of core-collapse SNe. This includes the necessity of mixing in SNe Ib/c (e.g., Wheeler et al. 1987; Shigeyama et al. 1990; Hachisu et al. 1991) and the presence of instabilities, such as convective instabilities in neutrino-driven explosion mechanisms and Rayleigh-Taylor instabilities during the expansion of the ejecta (e.g., Kifonidis et al. 2000, and references therein). In fact, with a small core mass ($`34M_{\mathrm{}}`$), SN 1993J should have relatively large Rayleigh-Taylor instabilities that mix the core (Iwamoto et al. 1997).
The observation of small-scale structure in SN line emission requires a spectrum with a high signal-to-noise (S/N) ratio and the spectral resolution to distinguish narrow features. This in turn necessitates a bright source and an efficient detector with at least moderate dispersion ($`110`$ ร
/pixel). These two requirements for SNe were first fulfilled with SN 1985F. Analysis of spectra of the Type Ib SN 1985F revealed several distinct clumps in the \[O I\] $`\lambda \lambda `$6300, 6364 lines as well (Filippenko & Sargent 1989). These clumps had widths $`100250`$ km s<sup>-1</sup> with amplitudes of $`210`$%. The small-scale features changed slightly over time, becoming narrower and relatively stronger. The Mg I\] $`\lambda `$4571 also had features, but they did not match those of the oxygen lines.
The bright, nearby SN 1987A provided another opportunity to study SN spectra in detail. While examining the \[O I\] $`\lambda \lambda `$6300, 6364 doublet of SN 1987A, Stathakis et al. (1991) discovered what appeared to be a sawtooth profile on the tops of the lines: there were $``$ 50 small-scale features with a typical size of 80 km s<sup>-1</sup> (full width at half maximum \[FWHM\]) and amplitudes of $`310`$% of the global line profile. The structures persisted over time and were consistent between the two doublet lines. Clumps were also observed in the H$`\alpha `$ line (Hanuschik et al. 1993). These had widths of $`400500`$ km s<sup>-1</sup> with amplitudes of $`710`$% of the overall line strength. The H$`\alpha `$ clumps were present over several hundred days, but individual clumps were not persistent. Hanuschik et al. (1993) interpreted this as changes in the level of dust extinction along the lines of sight to the individual clumps. Spyromilio & Pinto (1991) found that the oxygen density derived from the line intensity ratios of \[O I\] $`\lambda `$6300 and \[O I\] $`\lambda `$6364 gave too large an oxygen mass if the oxygen filled the entire volume suggested by the expansion velocity; a low filling factor for oxygen was consistent with the calculated density, volume at that epoch, and a reasonable oxygen mass. Such a low filling factor can be interpreted as clumpiness. Further evidence for clumps in the ejecta of SN 1987A is presented by Spyromilio, Stathakis, & Meurer (1993), who found a common feature in emission lines of H$`\alpha `$, \[Fe II\] $`\lambda `$7155, and \[Ca II\] $`\lambda \lambda `$7291, 7324. Clumps were also seen at later times in infrared (IR) emission lines of iron in SN 1987A (Spyromilio, Meikle, & Allen 1990; Haas et al. 1990). The pattern seen in the oxygen lines was not replicated in these other species.
The Type II SN 1988A may also have had clumps. They were not apparent in the spectra, but following the analysis of the \[O I\] $`\lambda \lambda `$6300, 6364 doublet in SN 1987A of Spyromilio & Pinto (1991), Spyromilio (1991) derived a low filling factor for oxygen in SN 1988A of 0.05, and thus deduced a clumpy nature for the source of the oxygen emission. In addition, Fassia et al. (1998) invoked clumps of helium to explain the strength of He I $`\lambda `$10830 line in the Type II SN 1995V given their model of <sup>56</sup>Ni mixing, although the clumps were not evident from the spectra.
SN 1993J, as noted before, was extremely bright, so any clumps in its spectrum should be readily detectable. Li et al. (1994) reported clumps in their spectra of SN 1993J. Wang & Hu (1994) noted the presence of narrow components in the forbidden lines of oxygen, and they used the clumpy nature of the SN to explain apparent blueshifts in the emission lines of \[O I\] $`\lambda `$5577 and \[O I\] $`\lambda `$6300. Spyromilio (1994, hereinafter S94) presented a much more detailed analysis of the clumps visible in the spectra of SN 1993J and also concluded that there is large-scale anisotropy in the SN based on the blueshifts of the lines (oxygen and magnesium) as well as the differences in the distribution of clumps both between and within the observed atomic species. This interpretation of the blueshifts was questioned by Houck & Fransson (1995), however. They concluded that the blueshifts are only apparent, and did not represent physical asymmetries in the lines. The \[O I\] $`\lambda \lambda `$6300, 6364 doublet can be dramatically affected by scattering of H$`\alpha `$ if oxygen and hydrogen are distributed in shells, and this would mimic the observed asymmetry. Blending affects the other lines, with \[O I\] $`\lambda `$5577 contaminated by \[Fe II\] $`\lambda `$5536 and \[Co II\] $`\lambda `$5526. The Mg I\] $`\lambda `$4571 line is affected by Fe II multiplets near 4600 ร
.
### 2.2 Analysis of the Clumps
As the brightest supernova in the northern hemisphere since SN 1954A, SN 1993J provided an object for which high S/N ratio spectra could be obtained over an extended period of time. Any intrinsic small-scale structure in the emission lines of SNe is normally lost in the noise of the spectra. For SN 1993J, the noise over most of our range ($`35009000`$ ร
) does not obscure the small-scale features in the spectra until several hundred days after the explosion. This allows us to make an extensive study of line substructure as the supernova evolves. Telluric absorption could introduce spurious small-scale structure, but we attempt to remove its effects (see Appendix).
Careful inspection of the spectra of SN 1993J reveals small-scale structure in many emission lines, especially \[O I\] $`\lambda \lambda `$6300, 6364, \[O I\] $`\lambda `$5577, O I $`\lambda `$7774, Mg I\] $`\lambda `$4571, and \[Ca II\] $`\lambda \lambda `$7291, 7324. The \[O I\] $`\lambda \lambda `$6300, 6364 lines in particular show a distinctive sawtooth profile similar to those described above for SN 1987A and SN 1985F (Figure 1). Note that all spectra have had the systemic velocity of SN 1993J removed ($`140`$ km s<sup>-1</sup>, determined from narrow lines in early-time spectra; see, e.g., FMH93). To isolate these fluctuations, we removed the global shape of the underlying line by smoothing the profile, and then subtracting the smoothed version. The spectra were smoothed with a running boxcar with a width of 60 ร
โwide enough to remove the small-scale fluctuations from the line profile, but narrower than the line itself, so that the subtraction would leave only the small features. This is the same method employed by Filippenko & Sargent (1989) to isolate similar substructure in the \[O I\] $`\lambda \lambda `$6300, 6364 lines of SN 1985F, although they used a smaller boxcar width for smoothing that was better matched to the velocity width of the fluctuations in SN 1985F.
Such a technique is similar to the process of unsharp masking in Fourier space. We experimented with this and other methods in the Fourier domain to isolate high-frequency components of spectra. The results were not superior to the smoothing technique. Moreover, the Fourier techniques became less successful as the relative noise level increased in later spectra. The subtraction of a smoothed line from itself does, however, introduce an artifact in the resulting residuals due to poor removal of the edges of the global line profile. These artifacts have a different character than the legitimate fluctuations; they are broader and generally show a gradual decrease from the subtracted continuum. Once an actual feature is reached, it clearly stands out from the edge effect. This is illustrated in Figure 2 with an artificial profile from which the smoothing technique isolates the six introduced components. The edge effects are prominent, but distinctly different in velocity width and character from the actual fluctuations. To eliminate the edge effects would require a model of the underlying line profile. This model would vary from line to line, and would also change over time as different amounts of nearby lines contaminate the feature under consideration. For example, as illustrated in Figure 4 of FMB94, the relative levels of \[O I\] $`\lambda \lambda `$6300, 6364 and H$`\alpha `$ $`\lambda `$6563 can change dramatically over a time scale of months. Subtracting a different line profile for different spectra at different times could introduce spurious components or temporal differences. In the interest of consistency, we chose to use the smoothing technique to identify substructure in the lines. In no case did it fail to find components visible in the original spectra, and the only new features introduced are the edge effects that are easily distinguishable from genuine components.
Figure 3 shows a typical example of the clumps in the \[O I\] $`\lambda \lambda `$6300, 6364 lines on day 209. There are five distinctive maxima marked in the figure. S94 found six clumps, but his clump *f* (at $``$ 2900 km s<sup>-1</sup>) is difficult to discern in this doublet; his clumps *a, b, c, d,* and *e* correspond to our clumps 5, 4, 3, 2, and 1. For the \[O I\] $`\lambda `$6300 line, these clumps are at velocities of (starting from the blueshifted edge) $`3220`$, $`2340`$, $`1510`$, $`310`$, and 410 km s<sup>-1</sup> (where 6300 ร
defines zero velocity). The marked minima (identified as 6, 7, and 8) have velocities of $`2750`$, $`1930`$, and $`740`$ km s<sup>-1</sup>, respectively. The uncertainty in the values for the velocities of clumps is $``$ 50 km s<sup>-1</sup>. The corresponding locations of these components for the \[O I\] $`\lambda `$6364 line are also indicated, but the features themselves are difficult to distinguish. There is also an emission feature at $``$ 2000 km s<sup>-1</sup>, but it is unclear if this is from \[O I\] $`\lambda `$6300, \[O I\] $`\lambda `$6364, both, or H$`\alpha `$. In fact, the large velocity width of H$`\alpha `$ at these times (FWHM $``$ 17,000 km s<sup>-1</sup>, full width at zero intensity \[FWZI\] $``$ 23,000 km s<sup>-1</sup>, FMB94; see also below) significantly contaminates the spectrum to a blue wavelength of at least 6375 ร
, possibly as far as 6310 ร
. The presence of H$`\alpha `$ throughout the region of the spectrum where one would find \[O I\] $`\lambda `$6364 makes the identification of the \[O I\] $`\lambda `$6364 clumps problematic at best.
The \[O I\] $`\lambda \lambda `$6300, 6364 clumps described above first become obvious in our day 93 spectrum. They are not apparent on day 45. As can be seen in Figure 4, once the features appeared, they remain consistent until day 433. Beyond that day, the clumps may still be present, but the supernova had faded to the point that it was difficult to obtain spectra with a S/N ratio high enough to distinguish the clumps. The clumps at $`2340`$ and 410 km s<sup>-1</sup> are, in fact, not obvious even in the day 433 spectrum. There are some slight changes in the individual clumps. The clump at $`1510`$ km s<sup>-1</sup> appears to broaden slightly, from 370 km s<sup>-1</sup> on day 93 (all widths are FWHM) to 470 km s<sup>-1</sup> on day 167, while increasing in amplitude in comparison with the other clumps. The amplitude increase continues past day 167 to the time when it became difficult to distinguish the clumps. The clump at $`310`$ km s<sup>-1</sup> grows considerably in relative strength, by almost a factor of two from day 93 to day 433. The clump at 410 km s<sup>-1</sup> narrows from 540 km s<sup>-1</sup> on day 93 to 390 km s<sup>-1</sup> on day 167, after which it remains constant.
The features seen in the \[O I\] $`\lambda \lambda `$6300, 6364 lines are also found in other oxygen emission lines. Figure 5 compares \[O I\] $`\lambda \lambda `$6300, 6364 with \[O I\] $`\lambda `$5577 and O I $`\lambda `$7774. The clumps line up quite well between \[O I\] $`\lambda `$6300 and \[O I\] $`\lambda `$5577. Li et al. (1994) found a similar correspondence for a single epoch. The only significant differences are slight changes in relative intensity of some of the clumps (compare the clumps at $`310`$ and 410 km s<sup>-1</sup> in \[O I\] $`\lambda `$6300 and \[O I\] $`\lambda `$5577). These may suggest intrinsic differences in physical conditions in the two clumps, but are more likely the result of contamination of the lines by other species (see above and Houck & Fransson for a discussion of potential line contaminants in SN 1993J). S94โs clump *f* at $``$ 2900 km s<sup>-1</sup> is apparent in \[O I\] $`\lambda `$5577 and O I $`\lambda `$7774. The O I $`\lambda `$7774 line is missing the $`2340`$ km s<sup>-1</sup> component completely, and the $`3220`$ km s<sup>-1</sup> clump is actually at $`3110`$ km s<sup>-1</sup>, slightly redshifted in comparison with the other oxygen lines. The shift of a clump is most likely the result of contamination, but the absence of one clump from O I $`\lambda `$7774 that is clearly evident in \[O I\] $`\lambda `$6300 and \[O I\] $`\lambda `$5577 is puzzling. The O I $`\lambda `$7774 line is contaminated by telluric absorption (see Appendix), but this is not likely to result in the apparent loss of a single component, especially not consistently over many epochs. The difficulty in determining actual flux values for the clumps effectively precludes any evaluation of the differing physical parameters for the various clumps, but the changing physical conditions from clump to clump are most likely the cause of the slight differences seen when comparing the various oxygen lines. In addition, the contamination of all the lines makes even global evaluations highly uncertain.
There is evidence for clumps in lines of other species. Figure 6 shows the small-scale structure of Mg I\] $`\lambda `$4571 from day 139 to day 298, while Figure 7 shows them for \[Ca II\] $`\lambda \lambda `$7291, 7324 from day 123 to day 298 (7291 ร
defines the zero velocity for calcium). (The fact that small-scale fluctuations in magnesium and calcium lines are visible for only a subset of the number of days that they are present in oxygen is more indicative of the vagaries of observation than a difference in the physical conditions producing the lines; earlier spectra are contaminated by other lines, while later spectra are noisier.) Again, the features are fairly consistent over time, but, as Figure 8 shows in a detailed comparison of \[O I\] $`\lambda \lambda `$6300, 6364, Mg I\] $`\lambda `$4571, \[Ca II\] $`\lambda \lambda `$7291, 7324, and H$`\alpha `$ on day 209, the pattern in the clumps is very different between the species, although Li et al. (1994) found that some features did correspond between H$`\alpha `$ and \[O I\] $`\lambda `$6300. Figure 9 shows \[O I\] $`\lambda \lambda `$6300, 6364, H$`\alpha `$, and Mg I\] $`\lambda `$4571 from day 433, and the differences between the lines are even more apparent. The H$`\alpha `$ line will be discussed in more detail in ยง3. The Mg I\] $`\lambda `$4571 line does have one clump (410 km s<sup>-1</sup>) that lines up with oxygen, but the rest of the structure is very different, with relatively fewer fluctuations, as in SN 1985F (Filippenko & Sargent 1989). (The minimum at $`2740`$ km s<sup>-1</sup> in \[O I\] $`\lambda \lambda `$6300, 6364 marginally lines up with a minimum in Mg I\] $`\lambda `$4571, but it is probably coincidental.) This may indicate that the magnesium and oxygen emission arise from substantially different clumps.
The \[Ca II\] $`\lambda \lambda `$7291, 7324 lines have some substructure, but they are relatively smooth. Aside from the peaks of the two lines and the edge effects described above, there is very little small-scale structure that is consistent. There is a small peak at $``$ $`1700`$ km s<sup>-1</sup>, but no other significant evidence for persistent clumps. In addition, these lines coincide with a telluric water absorption feature (cf. Figure 16), so incomplete removal of the water band profile may contaminate this region.
The scale of the clumps is not hidden by the resolution of the observations. The low-dispersion spectra used here have typical resolutions of $`67`$ ร
(FWHM). At 6300 ร
, that corresponds to $``$ 300 km s<sup>-1</sup>. Comparison with the high-resolution HIRES spectrum illustrates that there is little, if any, substructure with scales smaller than this (see Figure 9 of Paper I). Virtually all the clumps that are apparent in the HIRES spectrum appear in the low-dispersion spectrum taken four days later. The clumps that appear with lower contrast in the low-dispersion spectrum, but are easily found using our smoothing technique, do stand out more clearly in the HIRES spectrum. The HIRES spectrum shows that we are not missing any of the significant details of the clumps in the low-resolution spectra.
### 2.3 Discussion of the Clumps
The contrast between the presence of clumps in the emission lines of oxygen (and other species) and the relative smoothness of the calcium emission lines implies that the sources of emission for the various species are distributed differently in the supernova ejecta. One possibility is that the oxygen emission arises in distinct clumps of material while the source of calcium emission is distributed uniformly throughout the ejecta. While attempting to model the \[O I\] $`\lambda \lambda `$6300, 6364 lines and the Ca II lines (\[Ca II\] $`\lambda \lambda `$7291, 7324 and the near-infrared \[near-IR\] triplet), Li & McCray (1992, 1993) developed a similar concept to explain the structure seen in emission lines of SN 1987A that parallel those described above for SN 1993J.
Li & McCray (1992) initially focused on the forbidden oxygen lines. They developed a model based upon the relative line strengths at a given epoch for the \[O I\] $`\lambda \lambda `$6300, 6364 doublet that predicted a value for the total oxygen mass to filling-factor ratio. By choosing a reasonable value for the total oxygen mass (from theoretical models), they concluded that the oxygen emission must come from clumps distributed uniformly throughout the ejecta with a small filling factor. The clumps would develop from material synthesized in earlier evolution or during the explosion itself. As already mentioned, fast-moving oxygen components are seen in SNRs (Winkler & Kirshner 1985). When Li & McCray (1993) subsequently analyzed the calcium lines, they found that the observed lines were too weak given the probable amount of calcium produced during prior evolution and from explosive nucleosynthesis. Li & McCray hypothesized that clumps of calcium do form but do not contribute to the emission. These clumps of calcium intercept $`\gamma `$-ray radiation in proportion to their mass fraction. As this fraction is small, they receive only a small amount of radioactive luminosity and thus they do not achieve temperatures necessary for emission, although oxygen lines can be excited at these temperatures. The pre-existing envelope of the star is heated generally as hydrogen and helium do not radiate efficiently, allowing calcium that was thoroughly mixed into the atmosphere, most likely present from the time of the progenitorโs formation, to reach the temperatures required to produce Ca II emission.
This scenario of emission from smoothly distributed, pre-existing calcium while oxygen is clumped with a small filling factor explains the detailed line structure seen in SN 1993J. Other attempts to model the spectra of SN 1987A, however, concluded that the source of the oxygen emission consisted chiefly of the envelope, and so was also the result of pre-existing constituents (Swartz, Harkness, & Wheeler 1989; Swartz 1991), although later analysis by the same authors acknowledged that some oxygen from the core may contribute to the emission and that it might be clumped (Wheeler, Swartz, & Harkness 1993) . Smoothly distributed oxygen is difficult to reconcile with the clumps exhibited by the oxygen-line profiles. The underlying smooth line may be the result of the pre-existing oxygen, but clumps are clearly present in SN 1993J and SN 1987A, implying that there is a distinct difference in the site of oxygen and calcium emission.
The sawtooth appearance of the \[O I\] $`\lambda \lambda `$6300, 6364 lines indicates a rough scale for the size of the clumps in the ejecta. Chugai (1994) used similar data for SN 1987A to develop a technique for determining the oxygen mass. Given the minimum cloud size (on a velocity scale), the total velocity width of the lines, and the relative level of fluctuation in the clumps (the scale of the clumps compared to the overall line strength), one can then estimate the filling factor of the clumps ($`f0.1`$ in the case of SN 1987A). Several groups used the relative ratio of the \[O I\] $`\lambda \lambda `$6300, 6364 lines in SN 1987A to derive an oxygen mass to filling-factor ratio ($`M_O/f`$) of 11 $`M_{\mathrm{}}`$ (Li & McCray 1992) to 15 $`M_{\mathrm{}}`$ (Spyromilio & Pinto 1991; Chugai 1992; Andronova 1992). This implies a total oxygen mass of $`1.11.5`$ $`M_{\mathrm{}}`$ (Chugai 1994). As the line widths for SN 1993J are much larger (FWHM $``$ 5400 km s<sup>-1</sup> for SN 1993J on day 167 vs. FWHM $``$ 2700 km s<sup>-1</sup> for SN 1987A on day 173 \[Li & McCray 1992\]), the two doublet lines are blended, as well as severely affected by the broad H$`\alpha `$ line, and this effectively precludes a new evaluation for $`M_O/f`$. If one makes the admittedly dangerous assumption that this ratio is relatively constant among core-collapse SNe, then Chugaiโs method can be applied. For SN 1993J, the sizes of the clumps imply a typical cloud radius (in velocity space) of 150 km s<sup>-1</sup> with a relative fluctuation compared to the total line flux of $`\delta \mathrm{F}/\mathrm{F}0.1`$. To estimate the effective radius (in velocity space) for Chugaiโs calculation, we use Li & McCrayโs (1992) definition that emission at $`|v|v_{eff}`$ constitutes 90% of the flux in the line (Li & McCray refer to $`v_{eff}`$ as the โexpansion velocityโ). For SN 1993J, the FWHM is $``$ 5400 km s<sup>-1</sup>, and, assuming a Gaussian profile for the line, this implies that 90% of the emission of the line comes from within $`v\pm 3800`$ km s<sup>-1</sup>. With $`v_{eff}3800`$ km s<sup>-1</sup>, we find a filling factor of 0.06 by Chugaiโs method. Using the oxygen mass to filling-factor ratios derived for SN 1987A, these clump parameters yield an oxygen mass of $`0.70.9M_{\mathrm{}}`$ for SN 1993J.
The preferred models of Nomoto et al. (1993) and Shigeyama et al. (1990) for SN 1993J predict an oxygen mass of $``$ 0.4 $`M_{\mathrm{}}`$. This is similar to the value of $`0.47M_{\mathrm{}}`$ for oxygen in the model of Swartz et al. (1993). The models of Woosley et al. (1994) and Woosley, Langer, & Weaver (1995) yield $`0.50.7M_{\mathrm{}}`$ for SN 1993J. Utrobin (1994) finds a smaller value for the oxygen mass of $``$ 0.2 $`M_{\mathrm{}}`$, but also predicts a slightly narrower velocity profile for the oxygen lines ($`v4600`$ km s<sup>-1</sup>); the day 167 spectrum has a blue edge for the \[O I\] $`\lambda `$6300 line that indicates an expansion velocity of $``$ 4900 km s<sup>-1</sup>. For SN 1987K, another Type IIb SN, Schlegel & Kirshner (1989) found an oxygen mass of $``$ 0.3 $`M_{\mathrm{}}`$. The value we derive for SN 1993J is probably a bit high, but the blanket assumption that $`M_O/f`$ is constant among core-collapse SNe definitely increases the uncertainty of the result. A slightly smaller $`M_O/f`$ ratio would make our derived oxygen mass entirely consistent with the model predictions.
## 3 Late-Time Spectra
### 3.1 Late-Time Studies of Supernovae
Few SNe have had relatively frequent spectroscopic observations for more than three years. Several have have been recovered many years (even decades) after explosion. These include SN 1957D (Long, Blair, & Krzeminski 1989; Cappellaro, Danziger, & Turatto 1995), SN 1970G (Fesen 1993), SN 1978K (Ryder et al. 1993; Chugai, Danziger, & Della Valle 1995; Chu et al. 1999), SN 1979C (Fesen & Matonick 1993; Fesen et al. 1999, hereinafter F99), SN 1980K (Fesen & Becker 1990; Leibundgut et al. 1991; Uomoto 1991; Fesen & Matonick 1994; F99), SN 1985L (Fesen 1998), SN 1986E (Cappellaro et al. 1995), SN 1986J (Leibundgut et al. 1991; Uomoto 1991), and SN 1994aj (Benetti et al. 1998). SN 1988Z showed evidence for circumstellar interaction from its early-time spectra, and continues to be observed spectroscopically with $``$ 10 years of coverage (Filippenko 1991a, b; Stathakis & Sadler 1991; Turatto et al. 1993; Aretxaga et al. 1999). SN 1987A represents a special case; its proximity allows continued monitoring that would most likely not be possible for this SN if it were in a more distant galaxy, and so we will exclude it from this discussion<sup>1</sup><sup>1</sup>1As of early 2000, though, signs of more circumstellar interaction in SN 1987A are beginning to appear (e.g., Bouchet et al. 2000).
SNe such as SN 1988Z that show the spectroscopic characteristics of circumstellar interaction from an early time do so through very strong, narrow emission, especially in the Balmer lines. They are often referred to as a subclass of SNe II, either as โSeyfert 1โ types or the more commonly used Type IIn (Filippenko 1989, 1991a; Schlegel 1990). If we ignore the SNe IIn, and consider only the SNe II observed at very late times that appeared normal initially, then all of the SNe recovered after many years have fairly similar spectra. The pre-condition of normalcy at maximum brightness is not necessarily applicable to all the SNe listed above, as some (SN 1957D, SN 1978K, and SN 1986J) were not actually observed until much later in their evolution. All but SN 1978K have fairly broad emission lines (FWZI $``$ 10,000 km s<sup>-1</sup>) of H$`\alpha `$, \[O III\] $`\lambda \lambda `$4959, 5007, and \[O I\] $`\lambda \lambda `$6300, 6364 at late times. Most of these lines show a largely box-like profile, as expected from an expanding, roughly spherical, shell.
### 3.2 Prior Analysis of our Nebular to Late-Time Spectra
Spectra of SN 1993J up to day 433 are discussed by FMB94. They concluded that evidence for circumstellar interaction is visible as early as day 235 when the box-like profile of H$`\alpha `$ became distinct. This shape is predicted by the circumstellar interaction models of Chevalier & Fransson (1994; hereinafter CF94). Indeed, FMB94 found an H$`\alpha `$ luminosity of $`4.4\times 10^{38}`$ ergs s<sup>-1</sup> on day 433, in approximate agreement with the models of CF94 (and assuming the X-ray luminosity and density profile as described in Fransson, Lundqvist, & Chevalier ). There is a considerable amount of evidence for circumstellar interaction in SN 1993J from observations at radio and X-ray wavelengths (see Fransson et al. 1996, and references therein). Houck & Fransson (1996) and Patat, Chugai, & Mazzali (1995) also conclude that circumstellar interaction is present based on late-time optical spectra. The spectra described here are dominated by the effects of circumstellar interaction, and thus are discussed within that context. These late-time spectra are shown in their entirety in Figures 7 and 10 of Paper I.
For our present purposes, we restrict our definition of late-time spectra to those obtained from day 433 onward. Nebular features begin to appear by day 93 (cf. Figures 3, 5, and 7 of Paper I), and were noted earlier ($``$ day 62) by others with better temporal coverage during this period (Barbon et al. 1995). From day 209, the nebular features completely dominate the spectra, although an unusual box-shaped H$`\alpha `$ line may also be apparent at that point. This box-shaped line is clearly visible by day 355. From that day, the emission characteristics of the SN begin to be dominated by evidence for circumstellar interaction. Nebular features in the ejecta continue to play a significant role in the spectra, but are less important from day 473 onward. This section will focus on the late-time spectra during this interaction phase, from day 473 to day 2454. Figure 10 displays three high-quality Keck spectra of SN 1993J from days 976, 1766, and 2454 to illustrate the discussion of the late-time spectra.
### 3.3 Extinction
Analysis of the late-time spectra depends upon the amount of reddening that affects them. As absolute values for the line fluxes are difficult to obtain, line intensity ratios must suffice, and here the effects of reddening can be extreme. Various values for the extinction were found in photometric and spectroscopic studies of SN 1993J. Wheeler et al. (1993) used reddened blackbody models in comparison with early-time spectra to conclude that $`A_V=0.47\pm 0.06`$ mag (all reported values of $`A_V`$ assume $`R=A_V/E(\mathrm{BโV})=3.1`$). With similar techniques, Lewis et al. (1994) calculated $`A_V=0.58\pm 0.05`$ mag, while Clocchiatti et al. (1995) found $`A_V0.7`$ mag. Richmond et al. (1994) summarized several measurements of the extinction toward M81 in general and SN 1993J in particular. The average value is $`A_V0.6`$ mag, but ranges from $`A_V=0.2`$ mag to $`A_V=1.0`$ mag. Barbon et al. (1995) estimated $`A_V=0.9`$ from Na I D line equivalent widths. The Schlegel, Finkbeiner, & Davis (1998) dust maps show that the Galactic component is $`A_V=0.25`$ mag, not an insignificant fraction. For the purposes of this discussion, we will take the extinction toward SN 1993J to be $`A_V=0.6`$ mag. All fluxes and line ratios will be reported for the spectra as measured directly as well as corrected for this amount of reddening using the extinction corrections of Cardelli, Clayton, & Mathis (1989), including the OโDonnell (1994) modifications at blue wavelengths.
### 3.4 Line Measurements
The emission-line intensity ratios of the late-time spectra of SN 1993J are fairly difficult to measure. The large velocity width of the lines (FWHM $`15000`$ km s<sup>-1</sup>; see, e.g., Figures 7, 8, 10, 11, 12, and 13 of Paper I and below) and consequent overlapping make the deblending of the lines almost impossible. Without a clean, isolated line to provide a well-defined profile, all measurements will be affected by a relatively large uncertainty. Despite this problem, we were able to extract some information from the spectra.
The first step in the line measurement was the removal of the underlying โcontinuum.โ This is not an actual continuum from a photosphere, but rather a mixture of blended lines that combine to produce a relatively smooth pseudo-continuum, extending from the ultraviolet (UV) edge of our spectra to $``$ 5500 ร
. (The pseudo-continuum was removed over the entire range of our spectra; redward of 5500 ร
it is much weaker.) The circumstellar interaction itself can also produce a weak blue continuum (Fransson 1984). The stronger region of the pseudo-continuum is probably iron and iron-group element emission. Beyond 5500 ร
, these species also contribute, but there may be other sources responsible for the red light. As this is not a well-defined continuum, its removal is an additional factor in the uncertainty of any measurements. The subtraction of the pseudo-continuum was accomplished with a spline fit by hand to the spectra, but in a consistent manner for all of the spectra to be analyzed.
Line fluxes were calculated by summing the contributions from each pixel over a range determined by the best approximation of the edges of the line. In the cases of blended lines, total fluxes were measured. The edges of the lines provided the values for the maximum blue and red velocities for the given line(s). Table 1 lists the fluxes of the lines available in our spectra relative to H$`\alpha `$, with the second number indicating the line ratios when dereddened by $`A_V=0.6`$ mag. The relative spectrophotometry of the spectra is excellent (see Paper I for discussion of the observing and calibration details), so the major sources of uncertainty in the fluxes are line blending, choice of line boundaries, and the subtraction of the continuum. The relative flux values probably have uncertainties of $``$ 10%; individual measurements with larger uncertainties are denoted in Table 1. Table 2 reports the velocities of the lines determined from the same measurements; similar caveats about the uncertainties of the fluxes apply to the velocities, although the values for FWHM are more secure than either the blue velocity at zero intensity (BVZI) or the red velocity at zero intensity (RVZI).
For the day 976 spectrum (see Figure 10), we felt that the line shapes were sufficiently simple and well-defined compared to the other epochs to attempt a decomposition of the He I $`\lambda `$5876 + Na I D blend. We modeled the line as a simple box shape with an appropriate velocity width (15,000 km s<sup>-1</sup>). Using the SPECFIT task in the STSDAS package of IRAF, we found a good fit to the blend with approximately equal contributions from both He I $`\lambda `$5876 and Na I D (Figure 11). Later spectra were either too noisy or did not show as clean a box shape for the profile as did the day 976 spectrum. Visual inspection of the later spectra, however, indicates that the relative ratio of these two lines does not appear to change significantly. This can be seen in Figure 10 of Paper I. The blue edge of He I $`\lambda `$5876 and the red edge of Na I D are not as sharp in later spectra as they are on day 976. For days 2028 to 2176, the lower S/N ratio is at least partly responsible for degrading the line profile. By day 2454, though, the edges still appear, but the shoulders of the lines are more rounded. This may indicate that the spherical structure that produces the box-like profiles is beginning to lose its coherence. When comparing the relative heights of the blue and red shoulders of the He I $`\lambda `$5876 + Na I D blend, one must consider that the pseudo-continuum affects the blue half more than the red. Taking this into account, the relative strengths of the two lines appear to remain effectively constant. There may be a slight increase of the strength of He I $`\lambda `$5876 in comparison with Na I D, but it is not a significant change. Therefore, the line ratio for the He I $`\lambda `$5876 + Na I D blend reported in Table 1 can probably be interpreted as approximately half He I $`\lambda `$5876 and half Na I D.
We also employed a decomposition technique that used one of the lines in the spectra to model the others. The individual species were different enough that we chose to use the line of a given element to isolate only lines from the same species, although different ionization states were allowed. Given the lines available, this restricted us to the use of H$`\alpha `$ to model H$`\beta `$, and \[O II\] $`\lambda \lambda `$7319, 7330 to model \[O I\] $`\lambda \lambda `$6300, 6364 and \[O III\] $`\lambda \lambda `$4959, 5007. The line used as the model was isolated from the spectrum, shifted in wavelength space to match the other line, and then scaled until the subtraction of the model gave the most effective removal of the other line. The choice of scaling was subjective, as contamination by other species and noisy spectra precluded a more formal comparison. (The \[O I\] $`\lambda \lambda `$6300, 6364 doublet was removed from the blue edge of H$`\alpha `$ before it was used.) The fact that \[O II\] $`\lambda \lambda `$7319, 7330 matched the other oxygen lines so well suggests that there is very little, if any, contamination by \[Ca II\] $`\lambda \lambda `$7291, 7324 in this wavelength region at late times. Calcium is probably the more significant line at earlier times, as discussed in ยง2. The relative line fluxes determined with this technique are reported in Table 3. The line-subtraction technique was only effective in the spectra from day 976 onward. In addition, the results were more accurate for the high S/N ratio spectra from Keck (days 976, 1766, and 2454) than for the more noisy spectra from Lick (days 2028, 2069, 2115, and 2176).
### 3.5 Discussion
Most of our interpretation of the late-time spectra of SN 1993J falls within a comparison with the circumstellar interaction model of CF94. The CF94 model envisions cool, freely expanding SN ejecta colliding with circumstellar material from a pre-explosion stellar wind. A forward shock propagates into the wind, while a reverse shock moves back into the ejecta. The SN ejecta have a fairly steep density gradient, leading to a slow reverse shock with emission at far-UV wavelengths (possibly in X-rays, with a different gradient). This produces emission from highly ionized species. Absorption by a shell formed at the shock boundary can yield low-ionization lines, although these can also originate in the ejecta themselves. CF94 treat two different models for the structure of the wind. One is a power law, most applicable to a relatively compact progenitor, while the other uses the density structure of a red supergiant (RSG) from stellar evolution models. They make specific predictions for their model, including line intensity ratios and line profiles.
#### 3.5.1 Velocities
The uncertainty of the edges of the lines at zero intensity, as well as the large number of potentially overlapping lines, makes interpretation of the measured velocities problematic. One of the very specific predictions of CF94โs interaction model is that the line widths decrease with time. A pulsar-powered model would have velocity widths that *increase* slowly with time, and would have $`v1000`$ km s<sup>-1</sup> (Chevalier & Fransson 1992). The velocities listed in Table 2 appear to suggest a general decrease over time. The H$`\alpha `$ BVZI is severely contaminated by \[O I\] $`\lambda \lambda `$6300, 6364 at early times, and this contamination never completely vanishes. In addition, H$`\alpha `$ may be affected on the red side by He I $`\lambda `$6678. Given that He I $`\lambda `$5876 is approximately one-third the strength of H$`\alpha `$ (cf. Table 1 and ยง3.4), He I $`\lambda `$6678 probably only has a small effect on H$`\alpha `$, unless unusually high densities are altering the typical He I line intensity ratios (e.g., Almog & Netzer 1989). The RVZI for \[O III\] $`\lambda \lambda `$4959, 5007 is difficult to isolate, and thus is highly uncertain. The BVZI for H$`\beta `$ appears to grow fairly dramatically at late times, but this is more likely the result of contamination by another species than any change in H$`\beta `$ itself. A comparison of H$`\alpha `$ and H$`\beta `$ in velocity space on day 1766 and day 2454 (Figure 12), though, shows that the blue edge of H$`\beta `$ does not look like H$`\alpha `$ at late times and is probably some other species. The CF94 model does predict a strong Mg I\] $`\lambda `$4571 line, but this feature near H$`\beta `$ is at the wrong wavelength to be magnesium.
The BVZI of He I $`\lambda `$5876 and the RVZI of Na I D are more likely to reflect the actual limits of velocity space occupied by their respective species. The most accurate velocity information comes from the FWHM of H$`\alpha `$. This is much less affected by errors in the subtraction of the pseudo-continuum, and there is less impact by weaker lines. There is still some contamination by \[O I\] $`\lambda \lambda `$6300, 6364, but the line-subtraction technique described above indicates that it widened H$`\alpha `$ minimally. Perhaps the best evidence that the FWHM of H$`\alpha `$ is fairly accurate is how remarkably constant the values are. Probably from day 881, and definitely from day 1766, the FWHM is $``$ 15,000 km s<sup>-1</sup>, within our measurement uncertainty. The decrease found by day 2454 appears to be real. If this is the case, then the line widths are decreasing in accordance with the CF94 model. Marcaide et al. (1997) also found an apparent deceleration in the ejecta using very long baseline interferometry.
The CF94 model has other predictions relevant to the observed velocities. One is that highly ionized lines should have different velocity widths than low-ionization lines. Unfortunately, we cannot test this predication, as the difference is only $``$ 830 km s<sup>-1</sup> in the power-law model for the wind structure. Another is the scale of the line-profile widths. In the power-law model, the half width at zero intensity (HWZI) is $``$ 5000 km s<sup>-1</sup>, while for the RSG model, the HWZI is $``$ 4200 km s<sup>-1</sup>. For the lines of SN 1993J, the HWZI is $``$ 7500 km s<sup>-1</sup>. This is not surprising as the structure of the progenitor of SN 1993J would not have been that of a typical RSG. It may indicate that the particular power law used is not applicable to SN 1993J either. The four SNe discussed by F99 (1986E, 1980K, 1979C, and 1970G) have expansion velocity widths (HWZI) on the order of 5500 km s<sup>-1</sup>, implying that they have higher-mass envelopes than SN 1993J. This is understandable, as most models for SN 1993J postulate that it had a low-mass ($`0.20.5M_{\mathrm{}}`$) hydrogen envelope, so a smaller mass implies a larger velocity for a given amount of input energy (see Paper I and references therein). In all five cases, though, the velocities are larger than predicted by any of CF94โs models. In addition, the mass-loss rate of SN 1993J as determined from radio observations (Van Dyk et al. 1994) is $`(215)\times 10^5M_{\mathrm{}}`$ yr<sup>-1</sup>, and this is comparable to the values for the other four SNe \[($`219)\times 10^5M_{\mathrm{}}`$ yr<sup>-1</sup>; F99, and references therein\].
#### 3.5.2 Emission Lines
Table 4 lists the predicted line fluxes from the CF94 model, both for the power-law wind structure and the RSG model. The measured fluxes from our data range from 1.8 to 6.7 years, thus providing a comparison with several epochs of the model. Table 4 only contains the lines with which we can make any comparisons. The limitations of the spectra in wavelength coverage and quality at red wavelengths restrict the set of usable lines. The blending of the lines also complicates some direct comparisons.
The spectra from days 653, 670, 881, and 976 are closest to the CF94 year-2 predictions. For these days, assuming Na I D is half the blend of He I $`\lambda `$5876 + Na I D as described above, the sodium line is nearly half the predicted strength. The \[O I\] $`\lambda \lambda `$6300, 6364 doublet is only a quarter of the predicted value and \[O II\] $`\lambda \lambda `$7319, 7330 is three times stronger than in the model, but the \[O III\] $`\lambda \lambda `$4959, 5007 strength may be comparable. If the line subtraction technique is correct, and H$`\beta `$ is $`0.20.3`$ of the H$`\beta `$ \+ \[O III\] blend, then the observed oxygen-line flux is similar to the predicted value. Unfortunately, the \[Ne III\] $`\lambda `$3968 line was not available in all our spectra. Its value on day 653 is three times larger than predicted, but this region has the most uncertainty due to potential line contamination and the decreasing quality of the spectra toward blue wavelengths.
The later spectra are closer in time to the year-5 model of CF94, but we will also consider the year-10 models. The Na I D strength is still weaker than predicted, but closer to calculated values, and would actually fit well with the prediction for year 2. The relative strength of the sodium line does increase, as is the general trend in the models. The RSG model actually predicts a quite low value for Na I D, in contrast to the power-law model. While continuing to gain in relative strength, \[O I\] $`\lambda \lambda `$6300, 6364 is still far below the predicted values for the power-law model. In this case, the oxygen doublet is in agreement with the RSG value at year 10. The biggest difference between the observations and the models may be the strength of \[O II\] $`\lambda \lambda `$7319, 7330. In the models, this feature barely contributes, while the observations show it to be a significant line, with $`7080\%`$ of the strength of H$`\alpha `$. Unfortunately, the CF94 RSG model does not report the predicted strength of \[O II\] $`\lambda \lambda `$7319, 7330. The RSG model predicts an even higher strength for \[O III\] $`\lambda \lambda `$4959, 5007 relative to H$`\alpha `$ than the power-law model, and this is not what is observed, so we will concentrate on the predictions of the power-law model.
As \[O II\] $`\lambda \lambda `$7319, 7330 is expected to originate in the ionized ejecta along with \[O III\] $`\lambda \lambda `$4959, 5007, it is unclear why the singly ionized oxygen lines should be so strong, when the doubly ionized lines are slightly weaker than the models indicate unless the ionization structure is fairly different from the models assumptions. The density may explain the ratios, as will be discussed below. Chevalier & Fransson (1989) show that \[O II\] $`\lambda \lambda `$7319, 7330 is stronger in models of typical late-time spectra of core-collapse SNe without circumstellar interaction. Unfortunately, \[Ca II\] $`\lambda \lambda `$7291, 7324 is predicted to be an order of magnitude stronger than the \[O II\] lines, so the circumstellar model is more appropriate, although the depletion of calcium onto dust grains formed as the ejecta cool may affect the \[Ca II\] $`\lambda \lambda `$7291, 7324 line strength (e.g., Kingdon, Ferland, & Feibelman 1995).
The \[Ne III\] $`\lambda `$3968 line is again uncertain in these very late-time spectra, but it appears to be significantly stronger than the models predict. This makes the issue of ionization structure even more complicated, as this high-ionization line also originates in the ejecta. If the relative strengths of \[O II\] and \[O III\] are the result of a lower ionization of the ejecta, then the strength of \[Ne III\] is difficult to explain. This may imply that there are also abundance issues that affect the relative line strengths. As SN 1993J was assumed to have lost most of its hydrogen envelope (see Paper I and references therein), perhaps the relative amount of hydrogen has been depleted in the ejecta compared to the abundances used in CF94โs models, which were computed for normal SNe II. This could explain the apparent excess of emission from the lines that originate in the ejecta (\[O II\], \[Ne III\]), while the lines that originate in the shell (Na I D, \[O I\]) seem too weak, as all the line ratios are relative to H$`\alpha `$. The strength of \[O III\] actually fits with the predictions, and thus does not necessarily support this scenario.
Some of the lines predicted by CF94 are notable by their absence from our spectra. These include \[C I\] $`\lambda `$8729, \[N II\] $`\lambda \lambda `$6548, 6583, \[O II\] $`\lambda \lambda `$3726, 3729, Mg I\] $`\lambda `$4571, and \[S III\] $`\lambda \lambda `$9069, 9532. The carbon line would not be particularly strong, but the near-IR sulfur lines would be comparable in strength to H$`\alpha `$. Neither of these features is apparent, although hints of the sulfur line may exist in the day 2176 spectrum (Figure 10 of Paper I), but clearly at a lower strength. The \[O II\] $`\lambda \lambda `$3726, 3729 line could be contributing to the \[Ne III\] $`\lambda `$3968 line, but the distinctive double-horned oxygen line profiles described below do not appear and the wavelengths do not match well. As discussed above, the Mg I\] $`\lambda `$4571 line might be appearing at the blue edge of H$`\beta `$, but the velocities are not correct. In addition, the magnesium line should be two-thirds of the strength of H$`\alpha `$, and that is obviously not the case. The nitrogen lines could be hidden within the H$`\alpha `$ structure, but there should then be a shoulder on the blue edge of H$`\alpha `$ caused by the $`\lambda `$6548 component. The predicted total \[N II\] $`\lambda \lambda `$6548, 6583 flux is about one-third that of H$`\alpha `$, and if the line ratios are not affected by density considerations, then the $`\lambda `$6548 line may be too weak to show up on the H$`\alpha `$ profile. As we show below, there is contamination on the blue side, but from \[O I\] $`\lambda \lambda `$6300, 6364.
Most of these apparently missing lines have one attribute in commonโa low critical density. For \[O II\] $`\lambda \lambda `$3726, 3729, \[N II\] $`\lambda \lambda `$6548, 6583, and \[S III\] $`\lambda \lambda `$9069, 9532, the critical densities are all less than $`10^6`$ cm<sup>-3</sup> for $`T10,000`$ K. (These, and subsequent values for the critical density and other line diagnostics, were calculated with the IRAF/STSDAS NEBULAR suite of tasks.) The \[C I\] $`\lambda `$8729 line has a high critical density ($`>10^7`$ cm<sup>-3</sup>), so its absence, along with the lack of Mg I\] $`\lambda `$4571, may be explained by an origin in the shell, whose lines appear to be weaker than those from the ejecta. The critical density for \[O II\] $`\lambda \lambda `$7319, 7330 at $`T10,000`$ K is $`5.7\times 10^6`$ cm<sup>-3</sup>, while it is only $`6.2\times 10^5`$ cm<sup>-3</sup> for \[O III\] $`\lambda \lambda `$4959, 5007. This may explain the relative weakness of \[O III\], if the absence of the other lines does imply a density $`10^6`$ cm<sup>-3</sup>.
The \[O III\] lines may be able to further constrain the density. The \[O III\] $`\lambda `$4363 and $`\lambda \lambda `$4959, 5007 lines are contaminated by H$`\gamma `$ and H$`\beta `$, respectively. This contamination is potentially significant from days 553 to 1766, but probably less than 20% thereafter (cf. Tables 1 and 3). The relative contamination of H$`\beta `$ would be greater, and this would tend to drive the \[O III\] ($`\lambda \lambda `$4959, 5007/$`\lambda `$4363) ratio lower as it dilutes \[O III\] $`\lambda \lambda `$4959, 5007 if removed properly. Residual H$`\beta `$ contamination could drive the \[O III\] ($`\lambda \lambda `$4959, 5007/$`\lambda `$4363) ratio higher. Without considering the effects of the hydrogen lines, though, the \[O III\] ($`\lambda \lambda `$4959, 5007/$`\lambda `$4363) ratio is already $``$ 1 at day 670, implying that, to within the uncertainties of our measurements, we can probably ignore the hydrogen without affecting our interpretation significantly. To get \[O III\] ($`\lambda \lambda `$4959, 5007/$`\lambda `$4363) as low as 1, the temperature must be at least 10,000 K, even at the large densities considered here. The electron density under these conditions would be $`n_e10^8`$ cm<sup>-3</sup>. By day 2454, the ratio is $``$ 3, implying a density of $`n_e10^7`$ cm<sup>-3</sup>. For higher temperatures, the required density drops. If $`T`$ 30,000 K, for example, $`n_e10^7`$ cm<sup>-3</sup> on day 670 and $`n_e2\times 10^6`$ cm<sup>-3</sup> by day 2454. No matter what the temperature is, the increase of the \[O III\] $`\lambda \lambda `$4959, 5007 to $`\lambda `$4363 ratio with time implies a decreasing density, as one would expect in the expanding ejecta. Given the discussion of critical densities above, a density $`10^6`$ cm<sup>-3</sup> with a temperature of $`T10,00030,000`$ K appears to be consistent with the line emission.
The relative ratios observed are, for the most part, consistent with values determined by F99 from other late-time spectra of core-collapse SNe. Both SN 1980K and SN 1979C have very similar line ratios when compared with SN 1993J. This includes the large amount of emission in \[O II\] $`\lambda \lambda `$7319, 7330. SN 1986E has a much weaker \[O II\] feature (see also Cappellaro et al. 1995; they identify the line at $``$ 7300 ร
as the \[Ca II\] $`\lambda \lambda `$7291, 7324 doublet). SN 1979C is the only one of these four to have a measured \[Ne III\] $`\lambda `$3968 line, and it is not as strong as in SN 1993J. The lines predicted by CF94 that are missing in SN 1993J are also missing in these SNe, suggesting that they, too, have a density higher than the critical density of these lines.
A few of the late-time spectra were observed on photometric nights. For these, we calculated the luminosity of the H$`\alpha `$ line using a distance to M81 of $`d=3.6`$ Mpc (Freedman et al. 1994). On days 553, 670, and 881, the H$`\alpha `$ luminosity was (in units of $`10^{38}`$ ergs s<sup>-1</sup>) 2.5, 3.0, and 4.8, respectively. By day 1766, it was 0.86, and 0.37 on day 2454. (To correct for an extinction of $`A_V=0.6`$ mag, the luminosities would be multiplied by 1.5.) Both the day 553 and day 670 spectra were observed through relatively narrow slits on nights with variable seeing, implying a larger uncertainty for the luminosities calculated for those observations. The day 881 spectrum was taken with a 4$`\mathrm{}`$ slit width; thus, the value of 4.8 $`\times 10^{38}`$ ergs s<sup>-1</sup> is probably a better indication of the H$`\alpha `$ luminosity during these days. Note that FMB94 found the luminosity of the H$`\alpha `$ line to be $`4.4\times 10^{38}`$ ergs s<sup>-1</sup> on day 433โa comparable value within our probable uncertainties. The day 1766 and day 2454 spectra were also taken through narrow slits, with the seeing comparable to the slit width. Therefore, the luminosities at these later times could be larger (or smaller) than measured, but most likely by not more than a factor of two, implying a real decrease in H$`\alpha `$ luminosity by such late times. CF94 predicted an H$`\alpha `$ luminosity of 0.45, 0.19, and 0.09 $`\times 10^{38}`$ ergs s<sup>-1</sup> at 2, 5, and 10 years, respectively, for the power-law model. The RSG model gave an H$`\alpha `$ luminosity of 0.48 $`\times 10^{38}`$ ergs s<sup>-1</sup> at 10 years. Patat et al. (1995) found an H$`\alpha `$ luminosity of $`1.3\times 10^{38}`$ ergs s<sup>-1</sup> (corrected for extinction of $`A_V=0.1`$ mag) on day 368 (their day 367) and they interpreted this level of emission as a requirement for circumstellar interaction; radioactive decay alone was not enough. Our values are consistent with the Patat et al. (1995) model that requires circumstellar interaction to power the emission. The luminosities we observe at late times for SN 1993J agree with the values reported for other SNe observed at similar times (see F99 for a table of values).
#### 3.5.3 Overall Comparison with the Circumstellar Interaction Model
The CF94 model makes many predictions for the spectrum resulting from circumstellar interaction. The comparison of these predictions with the late-time spectra of SN 1993J and the prior study of F99 indicates that the model is somewhat successful, but needs refinement. The velocity widths in SN 1993J and the four SNe of F99 are larger than the model predicts, especially larger than those of the RSG model, but they do follow the expected trend of a general decrease with time. The line intensity ratios of the RSG model seem fairly inconsistent with the observed values. The power-law model is slightly better, as the trends of the relative ratios over time are consistent. One exception to this is the prediction of a strengthening of H$`\beta `$ relative to H$`\alpha `$ as the interaction continues; we do not see this in SN 1993J (cf. Tables 3 and 4), although the density of the emitting material may affect this result. The trend of a decreasing luminosity for the H$`\alpha `$ line at late times is also consistent between the observations and the models, although the absolute values differ.
The power-law model predictions may not be so different from the observed results if density effects are considered. Many of the emission lines predicted by the model that have low critical densities are absent from the spectra of SN 1993J and the four SNe discussed by F99. SN 1993J may have a relatively lower hydrogen abundance, as shown by its transformation from Type II to Type IIb. This will increase the relative intensity of other lines. The fact that similar values are seen in the four SNe of F99, though, implies that the low-mass envelope of SN 1993J is not altering the circumstellar interaction significantly. A larger density for the emitting region associated with the circumstellar interaction may be the main explanation for the deviation of the line intensity ratios from the predictions of the CF94 models. The relative weakness of lines predicted to arise in the ejecta, as opposed to the shell, argues that a depleted level of hydrogen may have some impact on the line-intensity ratios.
#### 3.5.4 H$`\alpha `$ Line Structure
One other aspect of the late-time H$`\alpha `$ line is that it shows a persistent, narrow (unresolved, FWHM $``$ 250 km s<sup>-1</sup> on day 1766) feature at zero velocity (see Figure 9 for a detailed view on day 433). Careful inspection of Figure 10 shows that the line is present in all three of the high S/N ratio, late-time spectra. It is not clear if this is emission from flash-ionized circumstellar material around SN 1993J (as noted in early spectra, see Paper I and references therein), or merely an H II region along the line of sight. Finn et al. (1995) noted the presence of this feature and found it to be slightly variable in strength. They also show (their Figure 7) an image of the environment of SN 1993J taken through a narrow-band filter centered on H$`\alpha `$. There is faint H II region emission evident around the SN. Aldering, Humphreys, & Richmond (1994), while studying earlier images of M81 to analyze the progenitor of SN 1993J, found that the colors indicated more than one star, and perhaps a faint OB association, that could be the source of the narrow hydrogen emission. Finn et al. (1995) concluded that the feature was probably not directly connected with SN 1993J. If the emission is not associated with SN 1993J itself, then it cannot originate very far from the line of sight to SN 1993J. Data from the Keck telescope on day 1766 were taken through a 1$`\mathrm{}`$ slit with $``$ 1$`\mathrm{}`$ seeing (the extraction width was 4.$`\mathrm{}`$2). The narrow feature is still present, and thus is likely directly along the line of sight. The H$`\beta `$ line shows no evidence for this feature, but the S/N ratio of the spectra in this region would make identification difficult.
Another possible source for the narrow H$`\alpha `$ line is material ablated from a companion star. As discussed in Paper I, many models for SN 1993J postulate a companion to facilitate the loss of the hydrogen envelope from the progenitor. Chugai (1986), Livne, Tuchman, & Wheeler (1992), and Marietta, Burrows, & Fryxell (2000) have considered the effects of the interaction of a SN explosion with a companion. Their calculations indicated that material from a companion to an exploding white dwarf could be swept up in the ejecta with emission becoming visible several hundred days after the explosion. The predicted widths, though, range from a few hundred to over a thousand kilometers per second. Given the low velocity width of the observed late-time narrow H$`\alpha `$ line in SN 1993J, it is unlikely that ablated material from the companion star is the source.
As can be seen in Figure 10, there are small-scale features covering the flat top of the H$`\alpha `$ profile similar to the clumps described in ยง2 for the oxygen lines. In this case, though, the structures change over time. As noted by FMB94, they probably represent Rayleigh-Taylor instabilities in the hydrogen-emitting region behind the shock (Chevalier, Blondin, & Emmering 1992). The instabilities are enhanced if the shock is radiative (Chevalier & Blondin 1995), as Fransson et al. (1996) suggest for SN 1993J.
#### 3.5.5 Late-Time Oxygen Lines
The relative strengths of the various oxygen lines are shown by the ratios determined using the line-subtraction technique described in ยง3.4 and presented in Table 3. The \[O I\] $`\lambda \lambda `$6300, 6364 to \[O II\] $`\lambda \lambda `$7319, 7330 ratio is approximately constant. The values for the ratio of \[O III\] $`\lambda \lambda `$4959, 5007 to \[O II\] $`\lambda \lambda `$7319, 7330 are difficult to interpret given their apparently large and random range. The values from the Keck spectra (days 976, 1766, and 2454) are the most reliable, and, considering only these values, the relative strength of \[O III\] $`\lambda \lambda `$4959, 5007 appears to weaken slightly compared with \[O II\] $`\lambda \lambda `$7319, 7330 over the course of the observations. This may indicate a reduction in the strength of the interaction and thus a decrease in the ionization level within the ejecta. Given the uncertainty in these values, this is highly speculative.
A striking change occurs in the spectra of SN 1993J between days 976 and 1766. The oxygen lines (\[O I\] $`\lambda \lambda `$6300, 6364, \[O II\] $`\lambda \lambda `$7319, 7330, and \[O III\] $`\lambda `$4363 and $`\lambda \lambda `$4959, 5007) all develop narrower profiles on top of the underlying box-like shape (cf. Figure 10). These features are not contamination by other lines, as they appear with the same relative velocities for all of the oxygen lines accessible in our spectra (Figure 13). The peaks of the narrower profiles are separated by $``$ 7100 km s<sup>-1</sup>, so these are not the two components of the oxygen doublets (the maximum velocity separation for the oxygen doublets is $``$ 3000 km s<sup>-1</sup> for \[O I\] $`\lambda \lambda `$6300, 6364). We believe that these profiles represent the blue and red peaks of a double-horned profile due to the ejecta colliding with a disk-like, or at least a somewhat flattened, region. (For a derivation of the two-horned profile from an expanding disk-like structure, see, e.g., Leonard et al. 2000; Gerardy et al. 2000.) The disk-like structure may be due to the interaction of circumstellar material with a binary companion, thought to be present according to models for mass loss from the progenitor of SN 1993J (see Paper I and references therein). The red horn is weaker than the blue horn due to differential extinction; we see the former through a much longer path in the ejecta. The large gap in our coverage between days 976 and 1766 means that the onset of the disk features was not recorded. There are vague hints of the incipient two-horned structure on day 976 (Figure 14), but it is certainly not as significant as in the later spectra. This type of structure is not evident in any of the four SNe whose late-time spectra are discussed by F99.
Starting with the day 1766 spectrum, a notch develops in the middle of the H$`\alpha `$ line. This loss of flux at zero velocity may indicate a slight flattening of the spherical distribution that had been the source of the hydrogen emission. Blondin, Lundqvist, & Chevalier (1996) present a model for an axisymmetric circumstellar interaction in which the SN is interacting with a dense equatorial structure. Such a scenario would result in little emission along the axis of symmetry because of a lower abundance of circumstellar material, while the equatorial structure would emit as in a typical interaction. This could create a disk-like or flattened source of circumstellar emission as is seen here. As the ejecta expand, they can interact with the disk-like or flattened circumstellar material producing the two-horned profiles. The emission-line profile from an expanding disk is double peaked and resembles that of a bipolar ejection. In fact, some new models for the core-collapse mechanism may require a bipolar outflow (Khokhlov et al. 1999), although they are fairly speculative. In addition, the optical spectropolarimetric studies of SN 1993J indicated asymmetries at early times (Trammell, Hines, & Wheeler 1993; Tran et al. 1997). Polarimetry of other SNe (e.g., Leonard et al. 2000; Wang et al. 2000) also suggests that asymmetries might be common, although the late appearance of a non-spherical emission source may indicate that the circumstellar matter is the cause of the asymmetry.
The appearance of this disk-like structure may complicate the interpretation of the line-intensity ratios discussed above. There is still the box-like profile from a roughly spherical source, in addition to the flattened region. The discussion of line ratios assumed that the conditions were the same for the entire emitting region. It may be that the shell and the disk have distinctly different physical conditions, but such differences would be very difficult to separate in our spectra. The wings of the disk-like profile are obvious, but the precise decomposition of the line into disk and shell components would be model dependent. This is made even more complicated by the fact that the box-like profile shows no significant asymmetry, while the two-horned profile is clearly stronger in the blue component. This implies that the flattened region suffers from differential extinction, while the roughly spherical shell does not.
## 4 Conclusions
The detailed substructure of the ejecta revealed by the spectra from early to nebular times indicates that SN 1993J is a clumpy supernova. (For a discussion of substructure that is the result of telluric absorption, see the Appendix.) Both oxygen and magnesium lines show clumps. The oxygen clumps appear in both permitted and forbidden lines, with some slight variations. The calcium lines do not seem to have clumps. This is consistent with a scenario in which the observed oxygen emission originates in clumps of the newly synthesized material (either through prior evolution or explosive nucleosynthesis), while the calcium emission arises mostly from pre-existing material distributed uniformly throughout the envelope, following the models of Li & McCray (1992, 1993). Using the technique of Chugai (1994), we find a total oxygen mass based on the observed clumps of $`M_O0.70.9M_{\mathrm{}}`$. This value is very uncertain and it is larger than the values predicted by models of SN 1993J.
We also present late-time spectra of SN 1993J. The lines clearly exhibit circumstellar interaction through their box-like emission profiles. They indicate a gradual decrease in velocity width from day 433 to day 2454, with the FWHM of H$`\alpha `$ remaining relatively constant from day 976 to day 2176 at $``$ 15,000 km s<sup>-1</sup>, decreasing slightly by day 2454. The line flux for H$`\alpha `$ is approximately constant from day 433 to day 881. It decreases by about a factor of five by days 1766 and 2454. Other strong lines in the spectrum include \[O II\] $`\lambda \lambda `$7319, 7330, \[O III\] $`\lambda `$4363, \[O III\] $`\lambda \lambda `$4959, 5007, \[Ne III\] $`\lambda `$3968, He I $`\lambda `$5876 + Na I D (probably each of equal strength), H$`\beta `$, and \[O I\] $`\lambda \lambda `$6300, 6364. These lines, and their relative ratios, agree fairly well with other studies of interacting SNe at late times (e.g., F99). While the circumstellar interaction model of CF94 predicts some lines that do not appear, and that \[O II\] $`\lambda \lambda `$7319, 7330 should be weak, the model is somewhat consistent with the general line-intensity ratios, and the predicted evolution of the line ratios agrees with the observations. Our spectra, though, indicate that the density is probably higher than that used by CF94. This may be because SN 1993J is unusual, but it is inconsistent with both the power-law model that assumes a compact progenitor and the RSG model that has a more extended progenitor. Considering SN 1993J and the four SNe of F99, the power-law model appears to be a better match in its predictions, although a higher density may be required to reproduce accurately the observed line-intensity ratios.
The late-time oxygen lines imply a relatively high density ($`n_e10^710^8`$ cm<sup>-3</sup>) on day 670, decreasing to $`n_e10^610^7`$ cm<sup>-3</sup> by day 2454, assuming a temperature of 10,000 K. The low ratio \[O III\] ($`\lambda \lambda `$4959, 5007/$`\lambda `$4363) $``$ 1 at early times excludes temperatures much below 10,000 K. The increase of this ratio to $``$ 3 on day 2454 shows the general expansion of the oxygen-emitting region. The appearance of the two-horned profile on the oxygen lines by day 1766 exhibits the development of emission from a somewhat flattened, disk-like structure in the interaction region. This may indicate a pre-existing equatorial belt of circumstellar material around SN 1993J, possibly similar to that observed in SN 1987A (e.g., Panagia et al. 1996, and references therein), although the emission in SN 1993J is likely the result of the ejecta colliding with the equatorial belt to produce the emission.
This research was supported by NSF grants AST-9115174 and AST-9417213 to A.V.F as well as by NASA through grants GO-7434, GO-7821 and GO-8243 from the Space Telescope Science Institute, which is operated by AURA, Inc., under NASA contract NAS 5-26555. We are grateful to the staffs of the Lick and Keck Observatories for their assistance with the observations. The W. M. Keck Observatory is operated as a scientific partnership among the California Institute of Technology, the University of California, and NASA. The Observatory was made possible by the generous financial support of the W. M. Keck Foundation.
## Appendix A Telluric Absorption
One aspect of spectroscopy that is often overlooked in the study of SNe is the impact of weak telluric absorption, which can sometimes mimic intrinsic absorption features in SN spectra. Virtually all observers are familiar with the prominent A-band ($``$ 7600 ร
) and B-band ($``$ 6850 ร
) absorptions produced by molecular oxygen, and some make an attempt to remove them. There are also several strong water absorption bands in the near-IR region. Weaker water bands in the range $`50006700`$ ร
, however, are much less well known. Typical SN spectra are often so noisy that the weaker absorptions are not obvious. This is especially true at blue wavelengths where detectors are usually less efficient; for example, there are ozone absorption features in the region $`32003450`$ ร
(the Huggins bands; e.g., Schachter, Filippenko, & Kahn 1989; Schachter 1991, and references therein). An extremely weak absorption, the Chappuis ozone band, extends from 5000 ร
to 7000 ร
, but it is unlikely to affect SN spectra significantly (see, e.g., Kondratyev 1969). Even in spectra with a moderate S/N ratio, some of the weaker bands can appear at high airmass or high humidity. The fortuitous appearance of a few relatively bright SNe in the past decade, however, shows that these features will be present even at low airmass in the spectra of all objects.
Figure 15 displays spectra of three bright SNe (SN 1993J, SN 1994D, and SN 1999em) over the range $`63606700`$ ร
. Each appears to contain an absorption line at $``$ 6515 ร
, but the spectra are shown with *observed* wavelengths, and these three SNe have differing redshifts ($`140`$ km s<sup>-1</sup>, 850 km s<sup>-1</sup>, and 720 km s<sup>-1</sup>, respectively), clearly implying a local source for the absorption. Figure 16 is a plot of the telluric absorption spectrum as determined from solar observations obtained with the Fourier-transform spectrometer at the McMath-Pierce telescope at Kitt Peak National Observatory<sup>2</sup><sup>2</sup>2NSO/Kitt Peak FTS data used here were produced by NSF/NOAO. (Wallace, Hinkle, & Livingston 1993, 1998). This telluric absorption spectrum has had the amount of water absorption increased by a factor of 1.7 to match better the conditions at Lick Observatory (G. Marcy, 1999, personal communication), and it has been smoothed to correspond to a spectral dispersion of $``$ 2 ร
/pixel. The inset of Figure 16 shows the telluric absorption in the region $`62506670`$ ร
. Note that at this resolution a strong, narrow absorption appears at 6515 ร
within a broader water band.
The presence of the absorption at 6515 ร
and other, weaker features near H$`\alpha `$ indicates that great caution must be taken when interpreting small-scale fluctuations in spectra near this wavelength range. This telluric line was noted and discussed as a possibly intrinsic feature in SN 1993J by Wheeler & Filippenko (1996) and Finn et al. (1995), although neither group attributed any significance to it. Since even experienced spectroscopists can be fooled by this line, it is important to be reminded of the potential effects of weak telluric absorption. This is especially true in efforts to search for the presence of H$`\alpha `$ in spectra of SNe Ia. A bright SN Ia would be an obvious choice for such an analysis, but a brighter object is more likely to show the effects of telluric absorption and lend itself to misinterpretation. The presence or absence of hydrogen in emission or absorption in the spectra of SNe Ia could have tremendous significance in the evaluation of their proposed progenitor systems (see, e.g., Branch et al. 1995). One must therefore be very careful with SN data to avoid the pitfalls of telluric absorption. With the development of ever larger telescopes, this problem will extend to even fainter SNe. In addition, the (relatively) strong O<sub>2</sub> absorption near $`62606340`$ ร
can appear to be an intrinsic SN feature in spectra with a low S/N ratio.
It is possible to remove most of the effects of telluric absorption, especially at high spectral resolution. The division by the spectrum of a comparison star that is intrinsically featureless in the wavelength regions affected by absorption can correct object spectra fairly well (Wade & Horne 1988; Bessell 1999). It can sometimes be difficult to find a star that is sufficiently featureless at the wavelengths of interest. This technique requires a well-exposed star to minimize the addition of any noise to the object spectrum. It is preferable to have a comparison star that is observed at an airmass similar to the object, but it is possible to scale the comparison star to match. We use our own procedures to remove telluric absorption following the technique of Wade & Horne (1988)<sup>3</sup><sup>3</sup>3We note that the IRAF task TELLURIC divides the object by the comparison star properly, but does not scale the spectrum of the comparison star according to the method of Wade & Horne (1988). With both TELLURIC and Wade & Horne, the comparison star is exponentiated by a function of the ratio of the airmasses of the object and the comparison star. In TELLURIC, however, this function is the ratio of airmasses multiplied by a scale factor, while with Wade & Horne it is the ratio of the airmasses raised to a power. Wade & Horne choose a power as the strongest telluric lines are saturated and the equivalent widths will grow with approximately the square root of the airmass. The TELLURIC scaling of the ratio of the airmasses effectively assumes an optically thin atmospheric absorption. When the airmasses of object and comparison star are similar, this different treatment has little impact. With the Wade & Horne method, variation of the power depending on the source and level of saturation of the absorption may result in better removal. We are conducting tests in this area..
In the normal course of data reduction, we remove the stronger telluric absorptions, but we only include the regions of weaker bands when observing at high airmass (or sometimes when the humidity is high). This is to avoid adding undue noise in these regions when the weak bands are already hidden in the noise of the spectrum. High S/N ratio objects, though, will show the weak bands. When we make an extra effort to remove them, as shown in Figure 15 for the 6515 ร
line in SN 1999em, we achieve some success. There are still residuals at 6515 ร
; it appears unlikely that any telluric absorption removal scheme can be 100% effective, but it can be done to a fairly good level, as long as interpretation of the spectra includes consideration of its effects. Observers must make sure that only real lines are identified as such.
|
warning/0006/cond-mat0006213.html
|
ar5iv
|
text
|
# Thermal equation of state of tantalum
## I Introduction
We investigate from first-principles the thermal equation of state of body-centered cubic (bcc) tantalum, a group V transition metal, which is a useful high-pressure standard due to its high structural mechanical, thermal and chemical stability. Ta has a very high melting temperature, 3269 K at ambient pressure, and its bcc structure is stable for a large pressure range. Static diamond anvil cell experiments up to 174 GPa and full-potential linearized muffin tin orbital (LMTO) calculations up to 1 TPa conclude bcc phase of Ta is stable at these pressure range. Similarly, shock compression experiments showed melting at around 300 GPa, but no solid-solid phase transition.
## II Static equation of state
Firstly, we discuss the static high pressure properties of Ta, which we obtained from first principles by using the LAPW method . The $`5p`$,$`4f`$,$`5d`$ and $`6s`$ states were treated as band states, others are described as core electrons. We used both the local density approximation (LDA) and generalized gradient approximation (GGA) for the exchange-correlation potential. The Monkhorst-Pack special $`\stackrel{}{k}`$-point scheme with a 16x16x16 k-point mesh (140 k-points within the irreducible Brillouin zone of the bcc lattice) was used after convergence tests. The convergence parameter RK<sub>max</sub> was 9.0, and the muffin tin radii were 2.0 bohr, giving about 1800 plane-waves and 200 basis functions per atom at zero pressure. The total energy was computed for 20 different volumes from 62.5 to 164 bohr<sup>3</sup> (1 bohr=0.529177 ร
), and the energies were fit to the Vinet equation ,
$`E(V,T)`$ $`=`$ $`E_0(T)+{\displaystyle \frac{9K_0(T)V_0(T)}{\xi ^2}}\{1+\{\xi (1x)1\}`$ (2)
$`\times exp\{\xi (1x)\}\}`$
where $`E_0`$ and $`V_0`$ are the zero pressure equilibrium energy and volume respectively, $`x=(\frac{V}{V_0})^{\frac{1}{3}}`$ and $`\xi =\frac{3}{2}(K_0^{}1)`$ , $`K_0(T)`$ is the bulk modulus and $`K_0^{}(T)=(K(T)/P)_0`$. The subscript 0 alone throughout represents the standard state P=0. All equations of state here are for an isotherm or static (T=0) conditions, unless specified otherwise. Pressures were obtained analytically from
$$P(T,V)=\{\frac{3K_0(T)(1x)}{x^2}\}exp\{\xi (1x)\}.$$
(3)
The calculated equation of state is compared with experiments in Fig. 1. The LAPW GGA results are found to be more accurate than the LDA. The discrepancies are larger beween theory and experiment at high pressures; this may be due to strength effects in the experiments , since we find good agreement with the experimental Hugoniot to 400 GPa (discussed below). For the pseudopotential mixed-basis calculations (MBPP, discussed below), we find that the LDA agrees fairly well with the experiments, indicating compensating errors between the pseudopotential and the LDA. Since we are using a first-principles approach, and want to avoid ad hoc variations in procedure to get better agreement with experiments, we use the computational more accurate method with the least approximations, that is LAPW and GGA, rather than the MBPP with LDA, in spite of the fortuitously better agreement of the latter with the room temperature data.
The residuals between the calculated and fitted energies show large deviations for volumes less than 80 bohr<sup>3</sup> (Fig. 2). Note that the residuals are all small and not evident for the large energy scale shown in Fig. 7a. Other equation of state formulations, such as the extended Birch equation show the same trend . When the fit is restricted to volumes greater than 80 bohr<sup>3</sup>, or extending the Vinet equation by two more parameters related to the next two pressure derivative of bulk modulus, it is improved significantly. Hence, these large residuals are related to different high- and low-pressure behavior of Ta. The band structures and densities of states (Fig. 3) show a major reconfiguration of the Fermi surface. This electronic topological transition is the reason for this change in compression, and the behavior of the residuals for the fitted equations of state. This indicates that systematic deviations from simple equations of state can be used to find subtle phase transitions .
Spin-orbit interactions may be important for Ta. In order to test this, we included spin-orbit coupling by second variational treatment including 20-80 bands. In contrast to the fully relativistic LMTO results , we found only negligible effect on the equation of state, so our computations were done without spin-orbit.
## III Calculation of Thermal Properties
In order to compute the high temperature properties of Ta, we separated the Helmholtz free energy as :
$$F(V,T)=E_{static}(V)+F_{el}(V,T)+F_{vib}(V,T),$$
(4)
where $`E_{static}(V)`$ is the static zero temperature energy, $`F_{el}(V,T)`$ is the thermal free energy from electronic excitations, and $`F_{vib}(V,T)`$ is the vibrational contribution to the free energy. $`E_{static}(V)`$ and $`F_{el}(V,T)`$ were computed using the LAPW method with the GGA. $`F_{vib}(V,T)`$ was computed using the particle-in-a-cell (PIC) model with a mixed basis pseudopotential method, as described below. The main differences we find between LDA and GGA are in the energy versus compression, but we find only small differences in energy versus atomic displacements. Since LDA converges much faster than GGA, and is also faster per iteration cycle, we used LDA for the large supercell computations required for the vibrational contributions. Differences from using GGA for this are negligible.
The electronic part $`F_{el}(V,T)`$ of the free energy is:
$$F_{el}(V,T)=E_{el}(V,T)TS_{el}(V,T)$$
(5)
where $`E_{el}(V,T)`$ is the internal energy due to thermal electronic excitations,
$$S_{el}(V,T)=2k_B\underset{i}{}f_i\mathrm{ln}f_i+(1f_i)\mathrm{ln}(1f_i)$$
(6)
is the electronic entropy, and the Fermi-Dirac occupation $`f_i`$ is
$$f_i=\frac{1}{1+\mathrm{exp}(\frac{(ฯต_i\mu (T))}{k_BT})},$$
(7)
$`ฯต_i`$ are the eigenvalues, $`\mu `$ is the chemical potential, and $`k_B`$ is the Boltzmann constant. The vibrational free energy is given in terms of the partition function as
$$F_{vib}(V,T)=k_BT\mathrm{ln}Z.$$
(8)
The particle-in-a-cell model was used to calculate the partition function. In this model, the partition function is factored by neglecting atomic correlations. An atom is displaced in its Wigner-Seitz cell in the potential field of all the other atoms fixed at their equilibrium positions, i.e. the ideal, static lattice except for the wanderer atom. The partition function is simply a product of identical functions for all the atoms, involving an integral of Boltzmann factor over the position of a single atom inside the Wigner-Seitz cell,
$$Z_{cell}=\lambda ^{3N}\{\mathrm{exp}\left[\frac{(U(\stackrel{}{r})U_0(T_0))}{k_BT}\right]๐\stackrel{}{r}\}^N,$$
(9)
where $`\lambda =h/(2\pi mk_BT)^{1/2}`$ is the de Broglie wavelengths of atoms and $`U(\stackrel{}{r})`$ is the potential energy of the system with the wanderer atom displaced by radius vector $`\stackrel{}{r}`$ from its equilibrium position. The advantage of the cell model over lattice dynamics based on the quasiharmonic approximation is that anharmonic contributions from the potential-energy of the system have been included exactly without a perturbation expansion. On the other hand, since the interatomic correlations between the motions of different atoms is ignored, it is only valid at temperatures above the Debye temperature. Diffusion and vacancy formation are also ignored, so premelting effects are not included. We have used the classical partition function, so quantum phonon effects are not included. Thus the heat capacity and thermal expansivity do not vanish at low temperatures, for example. The present results are appropriate for temperatures above the Debye temperature (245 K in Ta ) and below premelting effects. Since the Debye temperature is below room temperature in Ta, the classical thermal properties should be reasonable even down to room temperature.
The electronic thermal free energy was obtained using the Mermin theorem (Eq. 6). The charge density is temperature dependent through both occupation numbers according to the Fermi-Dirac distribution and self-consistency. The electronic contributions to the thermal free energy were computed by the LAPW method using the same computational parameters as the $`T=0K`$ computations described in Section II.
For the vibrational contributions, it is necessary to do a large number of large supercell calculations, which is computationally intractable by the LAPW method, but is achievable with the MBPP method . In this mixed-basis approach, pseudo-atomic orbitals and a few low-energy plane waves are used as the basis set within a density functional, pseudopotential calculation. It was shown that the method offers a computationally efficient but accurate alternative.
A semi-relativistic, nonlocal and norm-conserving Troullier-Martins pseudopotential (with associated pseudo-atomic orbitals) was used to describe the Ta atoms. The pseudopotential was generated from an $`5d^36s^26p^0`$ atomic configuration with cutoff radii 1.46, 2.6 and 3.4 bohr for $`5d`$, $`6s`$, and $`6p`$ potentials, respectively, with non-linear core corrections. The cutoff radii were optimized by testing the transferability of the pseudopotential by considering the reasonable variations of the reference atomic configuration and by comparing the logarithmic derivative of the pseudo-wavefunctions with all-electron values in the valence energy range. The $`6s`$ potential was chosen as the local component while $`5d`$ and $`6p`$ were kept as nonlocal while transforming the potential to the nonlocal separable Kleinman-Bylander form . A full plane-wave representation is used for the charge density and potential, and a smaller cutoff is used for the basis set. The pseudoatomic orbitals are expanded in the large plane wave set for evaluation of the potential and charge density integrals in the Hamiltonian and overlap matrices, and in the total energies. After checking the energy convergence, 550 eV and 60 eV were used for the large and small energy cutoffs, respectively, in the solid calculations. The exchange-correlation effects of electrons were treated within LDA. The $`T=0K`$ equation of state of bcc Ta was computed to test the pseudopotential, and is compared with experiment and LAPW results in Fig. 1.
For the PIC computations, a supercell with 54 atoms was used. The MBPP calculations were carried out on this 54 atoms supercell using LDA for exchange-correlations effects and 4 special $`\stackrel{}{k}`$ points for BZ integrations. The potential energy surface was then calculated as a function of the displacements of the wanderer atom. Symmetry was taken into account in order to reduce the number of computations. The integrand in Eq. 9 has a Gaussian-like shape and decays rapidly, and essentially is zero at half of the interatomic distances even at very high temperatures. Therefore, integration over the Wigner-Seitz cell can be replaced by an integration over the inscribed sphere. Also, the radial part of the integrand is invariant under point group operations of the lattice, hence a numerical quadrature can be used for angular integration based on the method of special directions . In this method, the radial integral is expanded in terms of lattice harmonics, cubic harmonics for a cubic lattice, then a quadrature rule is derived for the angular integration in terms of the radial integration by choosing special directions, $`\widehat{r}_i`$, in such a way that the contribution from $`l0`$ terms, as many lattice harmonics as possible, is zero. In all of the computations, we used one special direction which integrates exactly up to $`l=6`$ cubic harmonics . Then, the potential energy was calculated at 4-6 different displacements along this special direction. In order to model the potential, these computed values were fit to an even polynomial up to order 8, which shows the anharmonicity very clearly, since a second order fit describes the data poorly. Finally, the cell-model partition was calculated from Eq. 9 by carrying out the integration numerically, and the vibrational free energy is simply given in terms of partition function by Eq. 8.
## IV Thermal equation of state
We treated the resulting free energies three different ways. Firstly, F-V isotherms were fit using the Vinet equation of state, giving $`E_0(T)`$, $`V_0(T)`$, $`K_0(T)`$ and $`K_0^{}(T)`$ as the parameters of the fit (Table I, Fig. 4). Because of the thermal expansivity, the minimum energy shifts to higher volumes with increasing temperature, and above T=6000 K the P=0 volume is not in the range of volumes we studied. The experimental P=0 melting temperature is 3270 K, so temperatures above this are non-physical in any case. The parameters for higher temperatures are fictive parameters that describe the higher pressure equation of state accurately. The Vinet parameters could be fit to polynomials versus temperature to obtain a thermal equation of state, but the following approaches require fewer parameters.
A second thermal equation of state was obtained by analyzing the thermal pressure obtained from the Vinet fits (i.e. the differences in pressures between isotherms). The thermal pressure as a function of volume and temperature is shown in Fig. 5. The volume dependence of the thermal pressure for Ta is very weak up to 80% compression. Deviations at higher compressions may be due to the fit through the electronic topological transition discussed above, and thus due to inflexibility in the Vinet equation, rather than a real rise in thermal pressure. The thermal pressure is also quite linear in T. The thermal pressure changes are given by
$$P(V,T)P(V,T_0)=_{T_0}^T\alpha K_T๐T$$
(10)
and $`\alpha K_T`$ is quite constant for many materials in the classical regime (above the Debye temperature). Hence, a greatly simplified thermal equation of state is to add a thermal pressure
$$P_{th}(T)=aT$$
(11)
to the static pressure $`P_{static}`$. Since our equation of state is classical, we can use this expression at all temperatures. This simple equation of state gives $`q=\left(\frac{\mathrm{ln}\gamma }{\mathrm{ln}V}\right)_T=1`$ and $`\left(\frac{K_T}{T}\right)_V=0,`$ which therefore are good approximations over this pressure and temperature range for Ta.
The thermal pressure was averaged over volumes from 60 bohr<sup>3</sup> to 220 bohr<sup>3</sup>, and is shown as a function of temperature in Fig. 6. The solid line has a slope $`a`$ of 0.00442 GPa/K. This is close to $`\alpha K_{0T}(T)`$, which is 0.00460 GPa/K at 1000 K and zero pressure. So a simple equation of state for Ta is the static pressure given by V<sub>0</sub>(T=0)=123.632 bohr<sup>3</sup>, K<sub>0</sub>(T=0)=190.95 GPa, and K$`{}_{}{}^{}{}_{0}{}^{}`$(T=0)=3.98 in the Vinet equation (Eq. 3) plus the thermal pressure P$`{}_{th}{}^{}=0.00442T`$.
Thirdly, an accurate high temperature global equation of state was formed from the $`T=0K`$ Vinet isotherm and a volume dependent thermal free energy $`F_{th}`$ as:
$$F_{th}=\underset{i=1,j=0}{\overset{i=3,j=3}{}}A_{ij}T^iV^j3k_BT\mathrm{ln}T$$
(12)
This is the thermal Helmholtz free energy per atom, which must be added to equation 2 to obtain the total free energy. The parameters $`A_{ij}`$ are given in SI units in Table II, which gives the free energy in Joules/atom; T is in Kelvin and V in m<sup>3</sup>/atom. The term in $`T\mathrm{ln}T`$ is necessary to give the proper classical behavior at low temperatures, since we are evaluating the classical partition function, with $`C_V=3k_B`$ and $`S=\mathrm{}`$ at $`T=0K`$. For the best overall accuracy, the T=0 isotherm was also included in the global fit. The global fit is compared with the computed free energies in Fig. 7. The r.m.s. deviation of the fit is 0.4 mRy. At low temperatures (0 and 1000 K) the residuals (Fig. 7b) are larger due to the electronic topological transition discussed above (Fig. 2), but at higher temperatures this anomaly is less pronounced due to the thermal smearing, and the equation of state fits well.
Thermal equation of state parameters such as $`P`$,$`\alpha `$,$`\gamma `$,$`\delta _T`$,$`q`$, and the heat capacity $`C_V`$ and $`C_P`$ can be obtained from the global fit by differentiation and algebraic manipulation (see for a collection of useful formulas.) We now discuss the behavior of these parameters. The thermal expansion coefficient is presented, and compared with zero pressure experiments in Fig. 8. The deviations at lower temperatures are due to the use of the classical partition function. The thermal expansivity is a quite sensitive parameter, and the errors at moderate temperatures, which are typical in first-principles computations, may come from a number of sources (error in the P=0 volume, LDA, the pseudopotential, the PIC model, or convergence in k-points or basis set). The divergence in behavior at higher temperatures is not due to vacancy formation, since the vacancy formation in Ta is high (3.2 eV) , and the fraction of vacancies at the melting point is less than 10<sup>-4</sup>. The temperature range over which the anomaly occurs seems too large to be a premelting effect. One possible explanation would be an incipient solid-solid phase transition in Ta, which would not be detected in the PIC method. The upturn in $`\alpha `$ with increasing temperature is apparently a low pressure phenomenon. It is possible that it is due to an experimental problem, such as oxidation of the sample.
The thermal expansivity drops rapidly with increasing pressure (Fig. 8), and this is parametrized by the Anderson-Grรผneisen parameter (Fig. 9)
$$\delta _T=\left(\frac{\mathrm{ln}\alpha }{\mathrm{ln}V}\right)_T.$$
(13)
The behavior of $`\delta _T`$ is complex. At low pressures it increases with increasing temperature, but at elevated pressures it decreases with temperature. The parameter $`\delta _T`$ can be fit to a form $`\delta _T=\delta _T(\eta =1)\eta ^\kappa `$ where $`\eta =V/V_0(T_0)`$. The average $`\delta _T`$ (averaged from 0-6000 K) decreases with compression, and a power law fit gives $`\delta _T(\eta )=4.56\eta ^{1.29}`$; at 1000 K $`\delta _T=4.75\eta ^{1.17}`$. Interestingly these values are not that different from MgO ($`\delta _T(\eta =1,1000K))=5.00`$ and $`\kappa =1.48`$. The behavior is much different than for Fe, where $`\delta _T`$ is constant to 150 GPa with values of 5.2 and 5.0 for fcc and hcp, respectively, after which it drops more slowly than a power law . The difference between $`\delta _T`$ and $`K^{}`$ is an important anharmonic parameter, and is related to the change in the bulk modulus with temperature at constant volume and the thermal pressure with compression at constant temperature:
$`\delta _TK^{}`$ $`=`$ $`\left({\displaystyle \frac{\mathrm{ln}(\alpha K_T)}{\mathrm{ln}V}}\right)_T`$ (14)
$`=`$ $`{\displaystyle \frac{1}{\alpha K_T}}\left({\displaystyle \frac{K_T}{T}}\right)_V.`$ (15)
Fig. 9d shows that $`\delta _TK^{}`$ is quite small over a large temperature and compression range, but increases at high T and P. This is consistent with the accuracy of the simple equation of state (Eq. 11). The behavior of $`\delta _TK^{}`$ is also surprisingly similar to the behavior of MgO (see Fig. 3.3 in ).
Changes in thermal pressure $`P_{th}`$ are given by $`\alpha K_T=\left(\frac{P}{T}\right)_V`$ which is shown in Fig. 10. Changes in $`\alpha K_T`$ are small, but it is interesting that the sign of the change with temperature is strongly dependent on pressure, indicating that experiments to determine $`\left(\frac{\alpha K_T}{T}\right)`$ at low pressures may not be applicable to a very large pressure range.
A most important parameter, particularly for reduction of shock data, is the Grรผneisen parameter
$$\gamma =V\left(\frac{P}{E}\right)_V=\frac{\alpha K_TV}{C_V},$$
(16)
where $`E`$ is the internal energy. The Grรผneisen parameter is used in the Mie-Grรผneisen equation of state, which assumes $`\gamma `$ independent of temperature. Then the thermal pressure on the Hugoniot, for example is given by the change in internal energy by
$$P_{hug}P_{static}=\frac{\gamma }{V}(E_{hug}E_{static}).$$
(17)
Fig. 11 shows that at elevated pressures, $`\gamma `$ is moderately temperature dependent, and it varies more strongly with temperature below 100 GPa. The variation of $`\gamma `$ with pressure is given by
$$q=\frac{\mathrm{ln}\gamma }{\mathrm{ln}V}$$
(18)
which is shown in Fig. 12. The parameter $`q`$ is not constant, as is often assumed, but decreases significantly with pressure and temperature. If $`\alpha K_T`$ and $`C_V`$ were constant, Eq. 16 shows that $`q=1`$. Fig. 10 shows that $`\alpha K_T`$ is quite constant, so that large changes in $`q`$ must be due primarily to changes in the heat capacity $`C_V`$.
Fig. 13 shows indeed that the heat capacity is a strong function of temperature and pressure. This is due mainly to the electronic contributions. The experimental C<sub>P</sub> at zero pressure is also shown. Other than the large differences from experiment at very low temperatures, due to neglect of quantum phonon effects in the present model, there is a large increase in the experimental heat capacity with increasing temperature that is not seen in the PIC results. A similar large increase in the experimental thermal expansivity is not predicted by the model (Fig. 8). Vacancy formation (not included in the PIC model) seems an unlikely source for this discrepancy as discussed above. An incipient phase transition or sample oxidation seems the most likely cause of the observed behavior in the thermal expansivity and the heat capacity.
To compare with experiment at high pressures and temperatures, we consider the high temperature, high pressure equation of state obtained by shock compression . The pressures, $`P_H`$ and temperatures, $`T_H`$, on the Hugoniot of Ta are given by the Rankine-Hugoniot equation:
$$\frac{1}{2}P_H(V_0(T_0)V)=E_HE_0(T=0).$$
(19)
We solved the Rankine-Hugoniot equation using our equation of state results by varying the temperature at a given volume until it is satisfied. The calculated Hugoniot shows very good agreement with experimental data as seen in Fig. 14.
Computations of the Hugoniot using a modified free-volume method were recently presented by Wang et al. . Though superficially similar to the PIC method, their model is approximate. They included the electronic free energy in the same way that we do, but no supercell is used for the phonon contribution. Instead they find an effective potential from the equation of state of the primitive, one atom unit cell, and integrate the mean field phonon partition function based on this effective potential. This is a tremendous reduction in effort compared with the use of large supercells and finding the potential for displacing one atom in the supercell. They obtain impressive results for this simple model obtaining excellent agreement with the experimental Hugoniots, not only for Ta, but also Al, Cu, Mo and W. Nevertheless, it seems unlikely that this simply model will work for lower symmetry systems such as hcp-Fe or for elastic constants. For example, the c/a varies with temperature in hcp-Fe, but the Wang et al. model would not allow for this.
We summarize the zero pressure 300K equation of state parameters in Table III. The equation of state gives $`V`$(0,T=300 K)=124.489 bohr<sup>3</sup>, 2% higher than the experimental value. Another way of looking at the discrepancy, is that the computed pressure at the room temperature experimental volume 121.8 bohr<sup>3</sup> is 4.1 GPa rather than 0. Table III also shows the equation of state parameters computed at the experimental volume. The main discrepancy is the thermal expansivity which is 35% too high, though this is reduced by comparing at the experimental volume. The Gruneisen parameter is similarly high. The origin of this discrepancy is unknown, as discussed above, although the thermal expansivity (and thus $`\gamma `$) are known to be very sensitive. Apparently this inaccuracy must decrease with increasing pressure, since our Hugoniot agrees well with experiment, up to temperatures of almost 10,000K. Perhaps our potential surface is not modeled accurately enough at small displacements and low pressures due to the very small energy differences involved in that regime.
## V Conclusions
We have studied the static and thermal equation of state of Ta from first principles calculations. An electronic topological phase transition is found around 200 GPa. Three different forms of thermal equation of state is provided as: Vinet equation of state with temperature dependent equilibrium quantities, simple linear temperature dependent average thermal pressure, and a global fit to the Helmholtz free energy F(V,T). The simple equation of state $`P_{th}=aT`$ works quite well, but more accuracy and insites into higher order thermoelastic parameters were obtained from the global fit in $`V`$ and $`T`$. We find that $`\alpha K_T`$ is quite constant, as has been seen in experiments for a wide range of materials above the Debye temperatures and has been shown for simple pair potentials . Electronic excitations contribute significantly to the heat capacity temperature dependence of $`C_V`$, and thus to variations in the Grรผneisen parameter $`\gamma `$. We find good agreement with the experimental Hugoniot and thermal expansivity, though the rapid increase in the thermal expansivity and heat capacity at high temperatures remains unexplained.
###### Acknowledgements.
This work was supported by DOE ASCI/ASAP subcontract B341492 to Caltech DOE W-7405-ENG-48. Thanks to D. Singh and H. Krakauer for use of their LAPW code. Computations were performed on the Cray SV1 at the Geophysical Laboratory, supported by NSF grant EAR-9975753 and the W. M. Keck Foundation. We thank J.L. Martins, S. Mukherjee, G. Steinle-Neumann, L. Stixrude, and E. Wasserman for helpful discussions.
|
warning/0006/hep-ph0006131.html
|
ar5iv
|
text
|
# SEMI-INCLUSIVE VECTOR MESON PRODUCTION IN DIS
## Abstract
We analyze one-particle inclusive DIS in the case when a spin-1 hadron (such as a vector meson) is observed in the final state. We consider only leading order contributions in $`1/Q`$, but we include transverse momentum of partons. Several new fragmentation functions appear in cross sections. One of them can be measured in connection with the transverse-spin distribution function $`h_1`$.
The spin density matrix of a spin-1 particle can be decomposed on a Cartesian basis using the spin vector $`S^i`$ and a symmetric traceless rank-two spin tensor $`T^{ij}`$ :
$$๐=\frac{1}{3}\left(\mathrm{๐}+\frac{3}{2}S^i๐บ^i+3T^{ij}๐บ^{ij}\right),$$
(1)
where $`๐บ`$โs form a suitable basis of $`3\times 3`$ matrices. We parametrize the spin vector and tensor in the rest-frame of the hadron in the following way:
$$๐บ=(S_T^x,S_T^y,S_L),๐ป=\frac{1}{2}\left(\begin{array}{ccc}\frac{2}{3}S_{LL}+S_{TT}^{xx}& S_{TT}^{xy}& S_{LT}^x\\ S_{TT}^{xy}& \frac{2}{3}S_{LL}S_{TT}^{xx}& S_{LT}^y\\ S_{LT}^x& S_{LT}^y& \frac{4}{3}S_{LL}\end{array}\right).$$
(2)
The spin tensor of an outgoing vector meson can be extracted from the angular distribution of the decay products, although it is not possible to extract any information on the spin vector in an analogous manner. For instance, in the case of a $`\rho `$-meson decaying into $`\pi ^+\pi ^{}`$, the decay distribution depends only on the spin tensor components, being
$`W(\theta ,\phi )`$ $`=`$ $`{\displaystyle \frac{3}{8\pi }}\{{\displaystyle \frac{2}{3}}{\displaystyle \frac{2}{3}}S_{LL}(\mathrm{cos}^2\theta +\mathrm{cos}2\theta )\mathrm{sin}2\theta (S_{LT}^x\mathrm{cos}\phi +S_{LT}^y\mathrm{sin}\phi )`$ (3)
$`\mathrm{sin}^2\theta (S_{TT}^{xx}\mathrm{cos}2\phi +S_{TT}^{xy}\mathrm{sin}2\phi )\},`$
where $`\phi `$ is the azimutal angle between the meson production plane and the pions decay plane, and $`\theta `$ is the polar angle of one of the two pions in the $`\rho `$ rest-frame.
To describe semi-inclusive DIS we need to introduce two soft correlation functions, describing the quark distribution in the spin-1/2 target and the hadronization of a quark into the final state spin-1 hadron. In leading order in $`1/Q`$ (also referred to as โleading twistโ or โtwist-2โ) we are concerned only with the quark-quark correlation functions $`\mathrm{\Phi }`$ and $`\mathrm{\Delta }`$. The correlation function $`\mathrm{\Phi }`$ has been already widely studied in the literature (see e.g. ). The function $`\mathrm{\Delta }`$ is defined as (using Dirac indices $`\alpha `$ and $`\beta `$)
$`\mathrm{\Delta }_{\alpha \beta }(k,P_h,S_h,T_h)`$ $`=`$ $`{\displaystyle \frac{\mathrm{d}^4\xi }{(2\pi )^4}\mathrm{e}^{+\mathrm{i}k\xi }0|\psi _\alpha (\xi )|P_h,T_hP_h,T_h|\overline{\psi }_\beta (0)|0}.`$ (4)
Here, $`k`$ is the momentum of the quark decaying into an outgoing hadron after being struck by a virtual photon. The vector $`P_h`$ is the momentum of the outgoing hadron, $`S_h`$ is its spin vector and $`T_h`$ is its spin tensor. In the hadronic tensor we need the integrated correlation function
$`\mathrm{\Delta }(z,๐_T)`$ $`=`$ $`{\displaystyle \frac{1}{4z}}{\displaystyle dk^+\mathrm{\Delta }(k,P_h,S_h,T_h)}|_{k^{}=\frac{P_h^{}}{z};k_T=\frac{P_h}{z}},`$ (5)
which can be decomposed using 18 different fragmentation functions . If we do not observe the perpendicular component of the momentum of the outgoing hadron, we need to perform a further integration on $`๐_T`$ and to deal with the function
$`\mathrm{\Delta }(z)`$ $`=`$ $`{\displaystyle \frac{z}{4}}{\displaystyle \mathrm{d}^2๐_Tdk^+\mathrm{\Delta }(k,P_h,S_h,T_h)}|_{k^{}=\frac{P_h^{}}{z}}.`$ (6)
The parametrization of the correlation function after integration upon $`๐_T`$, as defined in Eq. (6), is
$`\mathrm{\Delta }(z)`$ $`=`$ $`{\displaystyle \frac{1}{4}}\{D_1(z)n/_{}+D_{1LL}(z)S_{hLL}n/_{}+G_1(z)S_{hL}\gamma _5n/_{}`$ (7)
$`+H_1(z)\mathrm{i}\sigma _{\mu \nu }\gamma _5n_{}^\mu S_{hT}^\nu +H_{1LT}(z)\mathrm{i}\sigma _{\mu \nu }\gamma _5n_{}^\mu ฯต_T^{\nu \rho }S_{hLT\rho }\},`$
where the fragmentation function $`H_{1LT}`$ is chiral odd and time-reversal odd. This function has already been discussed by Ji where it was named $`\widehat{h}_{\overline{1}}`$. Since $`\rho `$ decay into two pions does not allow to measure the spin vector, it is not possible to observe the fragmentation functions $`G_1`$ and $`H_1`$.
The cross-section of semi-inclusive deep-inelastic scattering is
$$\frac{\mathrm{d}\sigma (l+Hl^{}+h+X)}{\mathrm{d}x\mathrm{d}z\mathrm{d}y}=\frac{\pi \alpha ^2}{Q^4}\frac{y}{2z}L_{\mu \nu }\mathrm{\hspace{0.33em}2}MW^{\mu \nu },$$
(8)
where $`x`$, $`z`$ and $`y`$ are the usual scaling variables and $`L_{\mu \nu }`$ is the lepton tensor. The hadronic tensor $`W^{\mu \nu }`$ can be written as
$$2MW^{\mu \nu }=2z\text{Tr}\left[2\mathrm{\Phi }(x)\gamma ^\mu \mathrm{\hspace{0.33em}2}\mathrm{\Delta }(z)\gamma ^\nu \right].$$
(9)
By substituting the full structure of $`\mathrm{\Phi }`$ and $`\mathrm{\Delta }`$ into Eq. (9) and using the resulting hadronic tensor in Eq. (8), the cross section for an unpolarized target becomes:
$`{\displaystyle \frac{\mathrm{d}\sigma _U(l+\stackrel{}{H}l^{}+\stackrel{}{\stackrel{}{h}}+X)}{\mathrm{d}x\mathrm{d}z\mathrm{d}y}}=`$ (10)
$`{\displaystyle \frac{4\pi \alpha ^2s}{Q^4}}\left(1y{\displaystyle \frac{y^2}{2}}\right)xf_1(x)\left[D_1(z)+S_{hLL}D_{1LL}(z)\right],`$
while for a transversely polarized target we obtain:
$`{\displaystyle \frac{\mathrm{d}\sigma _T(l+\stackrel{}{H}l^{}+\stackrel{}{\stackrel{}{h}}+X)}{\mathrm{d}x\mathrm{d}z\mathrm{d}y}}=`$ (11)
$`{\displaystyle \frac{4\pi \alpha ^2s}{Q^4}}x(1y)|S_T||S_{hLT}|\mathrm{sin}(\varphi _{LT}+\varphi _T)h_1(x)H_{1LT}(z),`$
where $`\varphi _T`$ and $`\varphi _{LT}`$ are the azimuthal angles between the scattering plane and, respectively, the transverse spin of the target and the longitudinal-transverse spin of the outgoing hadron.
Eq. (11) shows that semi-inclusive DIS with spin-1 outgoing hadrons allows the measurement of the transverse-spin distribution function $`h_1(x)`$ in connection with the new fragmentation function $`H_{1LT}(z)`$. Investigation is required to evaluate the magnitude of this new function and of $`|S_{hLT}|`$. The measurement requires no perpendicular momenta to be recorded and no azimuthal asymmetries to be computed. It is sufficient, for instance, to reverse the spin direction of a transversely polarized target and calculate the relative single transverse spin asymmetry.
The complete list of cross sections involving transverse momentum dependent functions is also available .
|
warning/0006/astro-ph0006360.html
|
ar5iv
|
text
|
# SUNYAEV-ZELโDOVICH EFFECT REVIEW
## 1 General Presentation
The presence<sup>a</sup><sup>a</sup>ato be published in the Proceedings of the XXXVth Rencontres de Moriond, Energy Densities in the Universe, Editions Frontieres, 2000 of a hot teneous and fully ionised gas ($`T_e=10^8\mathrm{K}`$, $`n_e=10^3\mathrm{m}^3`$) in the intracluster medium was revealed with the first X-ray measurements toward clusters of galaxies. This gas which fell in the deep gravitational well of clusters of galaxies and thus heated up to very high temperature can only cool down via the free-free emission process. Only in the very center of clusters is the density of the electrons and nuclei enough for the cooling timescale to be less than the age of the Universe. Another cooling process exists via inverse Compton scattering on the (cold) cosmic microwave background.
This secondary cooling is called the Sunyaev-Zelโdovich (hereafter SZ) effect$`^\mathrm{?}`$. This effect preserves the number of CMB photons. If it were a pure scattering effect without energy change the CMB would not be globally affected. But there is a net energy gain by the CMB photons in the direction of clusters. The CMB is thus spectrally distorted. The adimensional Comptonisation parameter $`y`$ measures the SZ distortion:
$$y=\frac{kT_e}{m_ec^2}\sigma _TN_e,$$
(1)
where $`T_e`$, $`m_e`$, $`N_e=n_e๐l`$, and $`n_e`$ are resp. the electronic temperature, mass, column density and density. The integral is taken along the line of sight through the cluster. $`\sigma _T`$ is the Thomson cross-section. The second (and usually much weaker) SZ effect called kinetic effect is due to the peculiar cluster velocity $`v_c`$ and is measured by:
$$b=\frac{v_r}{c}\sigma _TN_e$$
(2)
Figure 1 shows the universal distortion spectrum produced by the thermal (dots) and kinetic (dashes) SZ effects.
The SZ effect is thus a radio, millimetre and submillimetre phenomenon. The thermal SZ effect has a very specific spectral signature (always negative for $`\lambda 1.4\mathrm{mm}`$) whereas the kinetic SZ effect is undistinguishable from the spectrum of the CMB primordial anisotropies. Both SZ effects are brightness effects which are spectrally independent of redshift in the observerโs frame, contrary to X-ray emission which shows the usual $`(1+z)^4`$ brightness dimming. A well-resolved cluster will show the same SZ effect whether it is at low or high redshift.
The energy density (which is the focus of this conference) of the CMB is enhanced towards clusters by the following amount:
$$E=4y\sigma _ST_{CMB}^4$$
(3)
The opacity $`\tau =\sigma _TN_e`$ and comptonisation parameter $`y`$ are of the order of a few $`10^2`$ and $`10^4`$ resp. in the richest clusters which therefore make the SZ effect a relatively small and linear distortion.
It has recently been acknowledged that one cannot neglect relativistic corrections to the (non-relativistic) universal spectral template, shown in Fig. 1, and independent on redshift. A complete review of the SZ effect is clearly beyond this presentation. An exhaustive recent review was made by Birkinshaw$`^\mathrm{?}`$.
## 2 Observational status
### 2.1 Astrophysical and Cosmological objectives
#### Witness the 3K remoteness
With several clusters firmly detected at a redshift up to about 0.5, the 3 K background cannot have an origin in the local Universe. There are not so many direct probes of the presence of the CMB at high redshift.
#### Measure the hot cluster gas distribution
The SZ effect directly measures the hot electron pressure along the line of sight. If one assumes a constant gas temperature, an electronic density following a King profile (with $`r`$ the radius from the cluster center, and $`a`$ the core radius) is often assumed:
$$n_e=n_{e0}(1+(r/a)^2)^{3\beta /2},$$
(4)
and will produce a SZ angular distribution
$$y_{SZ}=y_0(1+(\theta /\theta _a)^2)^{3\beta /2+1/2}.$$
(5)
The measurements are therefore linearly linked to the density of baryonic matter.
#### Total mass of gas
The quantity that is directly measured in a given experiment is really the brightness integrated over the instrument beam, something that can be loosely defined as $`\textcolor[rgb]{0,0,0}{Y}\textcolor[rgb]{0,0,0}{=}\textcolor[rgb]{0,0,0}{}_{\textcolor[rgb]{0,0,0}{b}\textcolor[rgb]{0,0,0}{e}\textcolor[rgb]{0,0,0}{a}\textcolor[rgb]{0,0,0}{m}}\textcolor[rgb]{0,0,0}{y}\textcolor[rgb]{0,0,0}{๐}\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}`$. The big virtue of SZ measurements is to give an easy access which is weakly dependent on redshift to the total gas mass of the cluster in the beam:
$$M_g2\times 10^{13}M_{_{}}(\frac{T_g}{10\mathrm{keV}})^1\frac{Y}{10^5\mathrm{arcmin}^2}f(z),$$
(6)
where the redshift function $`f`$ depends on the cosmological parameters (through the angular-diameter redshift dependence). The following table (computed for a critical standard model without cosmological constant) shows a subtle dependence of the measured SZ effect with redshift.
| $`z`$ | $`f\times 10^3`$ |
| --- | --- |
| 0.1 | 2 |
| 0.3 | 9 |
| 1.0 | 20 |
#### Total gravitational mass
Through the hydrostatic equation and from the gas pressure profile, one can deduce the total cluster mass $`M_T`$. Although this has been done with X-ray measurements, the SZ effect could in principle be used for a rather direct measurement of the total mass. It would be valuable, when precise measurements become available, to reassess the baryonic crisis: $`M_g/M_T\mathrm{\Omega }_b/\mathrm{\Omega }_0`$.
#### Peculiar radial velocity
The kinetic SZ effect is a 10 times weaker effect that the thermal effect. Accurate measurements of it in many cluster could in principle probe the large scale velocity field in the distant Universe. The CMB primary anistropies are in that case a โpollutionโ to these measurements which could be attempted by the Planck mission (Aghanim$`^\mathrm{?}`$). Ground-based attempts have so far provided upper limits (Holzapfel et al.$`^\mathrm{?}`$)
#### $`H_0`$ , $`q_0`$ measurement
The measurements of $`T_en_e๐l`$ with the SZ effect and $`n_e^2๐l`$ and $`T_e`$ with X-ray space observations yield an estimate of the true physical depth of the cluster. Assuming the cluster is spherical, this quantity can be compared with the angular size of the cluster and its redshift to give $`H_0`$. The weak dependence of that result on $`q_0`$ has been analysed, but the prospect of a serious measurement of it is marred by cluster evolution (see below). Measurements of the Hubble constant is clearly within reach, once systematic effects are well understood over a statistically significant sample of observed (local) clusters. The complementarity of the SZ effect with XMM-Newton and Chandra is obvious in that respect. This is one of the most important cosmological targets for SZ measurements. The second one is:
#### SZ Cluster number counts
Having SZ surveys over large area could provide a rather unbiased measurement of the cluster number counts. Optical and X-ray surveys have been known to be biased by chance alignments and resolution & surface brightness limits respectively. The weak dependence on redshift of the SZ effect (Eq. 6) makes a (costly) survey quite attractive for two reasons:
1. to determine the exact number density of local clusters (say of redshift less than 0.3) and the mass distribution function
2. to measure the evolution of this density and search for high redshift clusters.
#### Structured Matter Energy injection into the CMB
The average comptonisation parameter on any line of sight is about $`y10^6`$. From Eq. 3, the energy injection into the Universe by large scale non-linear structure formation is of the order of at most $`10^5`$ of the energy in the CMB itself.
#### Search for distant clusters : $`z>1`$
The search of SZ effect without the a-priori of X-ray maps have so far led to the as yet unconfirmed detection of two radio extended brightness decrements (Richards et al.$`^\mathrm{?}`$, Saunders et al.$`^\mathrm{?}`$). The secure detection of just few clusters at redshift above 1 would severely endanger models of the Universe with a critical matter density parameter (Bartlett et al.$`^\mathrm{?}`$).
### 2.2 The three observational techniques
#### Single dish radio telescopes
Pionneered by Birkinshaw and collaborators, this was the first standard technique to successfully measure the SZ effect. It was and remains the best adapted technique for local clusters with a very large extent (e.g. Coma: Herbig et al.$`^\mathrm{?}`$)
#### Radio interferometers
The most sensitive detections have been provided by radio interferometers which are less affected by sidelobe effects and can remove point sources by using the long baseline measurements. Moreover they benefit from large integration times (several months per year). The Ryle telescope (Saunders and Jones, this conference) in England and the OVRO-BIMA interferometer fitted (perverted) with radio receivers (Carlstrom et al.$`^\mathrm{?}`$) have obtained very sensitive maps of now tens of clusters with arcminute resolution at 15 and 30 GHz frequencies.
#### Bolometer photometers
The SuZie experiment provided the first detection of the SZ effect in the millimeter domain at the CSO 10m telescope (Wilbanks et al.$`^\mathrm{?}`$, Holzapfel et al.$`^\mathrm{?}`$). It is made of 6 bolometers at a wavelength of 2 mm. By using a drift scan technique whereby fixing the telescope in local coordinates, the cluster drifts through the detectors with the Earthโs diurnal motion, hence avoiding sidelobe effects and microphonic noise. The cluster A 2163 (the second X-ray brightest known cluster) was later detected at 1mm using the SuZie experiment with different filters and in the submillimeter domain by the balloon borne photometer PRONAOS-SPM (Lamarre et al.$`^\mathrm{?}`$) showing for the first time the change of sign of the SZ effect (see Fig. 1). Although in the (sub)millimeter domain, the point sources should not contaminate the SZ measurements so much, interstellar dust thermal emission must be dealt with to correct the measurements. Another limitation comes from sky noise: the water vapour is inhomogeneous in the atmosphere. Its emission in the telescope is variable in time and angle and frequency (see Fig. 1). The Diabolo experiment (Benoit et al.$`^\mathrm{?}`$) uses a dichroic beam splitter with six 0.1 K bolometers in order to simultaneously measure and hence subtract the water vapour emission at 1.2 mm (where the SZ effect is almost vanishing) from the SZ measurement at 2.1 mm. At the IRAM 30 m telescope, it provided the highest angular resolution of the SZ effect on several clusters with 30 arcsecond beam in 1995 (D sert et al.$`^\mathrm{?}`$) and 22 arcsecond beam since 1997 (Pointecouteau et al.$`^\mathrm{?}`$ and this conference).
## 3 The next SZ ten years
The sensitivity of present ground-based detectors is quite close to photon noise limits, typically an effective value $`y(1\sigma )10^4\mathrm{hour}^{\frac{1}{2}}`$ for a single bolometer. This is also typical for interferometers. All 3 kinds of observing techniques are also currently limited by the range of angular scales that can be measured, whereas the angular distribution of clusters of galaxies is widespread, from core radii to Abell radii, along with substructure scales. We can see that the main goals of SZ observations in the next years are:
1. Improve the statistics on $`H_0`$ measurements
2. Number counts of SZ clusters and high redshift cluster search
3. Detailed analysis of individual clusters
For point 1 and 2, interferometers are clearly very promising. Dedicated radio telescope arrays which are currently being built with modest size telescopes aim at covering a large range of angular scales (in particular the shortest baselines). By covering a large frequency bandwidth and by using smaller telescopes (!), improvements in detectivity could be better by as much as a factor of 1000 in the next 3 years (e.g. the AMI project: Jones, this conference). For point 3, bolometer arrays (with hundred to thousands of pixels) should bring a clear multiplex advantage over existing technology$`^\mathrm{?}`$. They will give unprecedented high angular resolution maps in the near future (20 arcseconds), that will be useful to study the detailed angular distribution of clusters (whether at low or high redshift) to unravel the cosmogony of large scale structures. In that respect, the structure of high redsihft clusters (e.g. Ebeling et al. $`^\mathrm{?}`$) which is far from smooth is interesting for cluster formation scenarios but may prove a show stopper for $`q_0`$ and other second order effect measurements.
Future space missions will provide a different perspective. Planck will give an unbiased catalog of at least several thousands of SZ sources. MAP, although not sensitive to individual clusters, may still see some signal by correlation with large scale structures as seen in the optical (Refregier et al.$`^\mathrm{?}`$).
The interpretation of SZ data is depending on the quality of other data, and vice-versa. The arrival of Chandra and XMM-Newton is a strong incentive for improving SZ measurements. Comparison with visible and near-infrared data obtained by large telescopes (substructures and weak lensing) is also very valuable.
We have clearly moved from detection experiments towards a powerful tool for the study of clusters. A global approach, using SZ observations but also other wavelengths, is a must for the understanding of clusters of galaxies.
## Acknowledgments
We wish to thank the organisers for such a pleasant and lively forum of discussions of which the SZ effect was one of the foci. We thank A. Refregier for allowing us to show the large scale SZ simulated map.
## References
|
warning/0006/hep-lat0006022.html
|
ar5iv
|
text
|
# GUTPA/00/06/02 Casimir scaling of SU(3) static potentials
## I Introduction
Non-perturbative QCD effects in general and the nature of the confinement mechanism in particular are theoretically challenging. At the same time these aspects are important for high energy and low energy particle and nuclear phenomenology. Several models of non-perturbative QCD have been proposed whose predictions happen to differ from each other substantially in some cases. Prominent examples are bag models , strong coupling and flux tube models , bosonic string models , the stochastic vacuum model , dual QCD , the Abelian Higgs model , and instanton based models . Lattice simulations of interactions between static colour sources offer an ideal environment for discriminating between different models of low energy QCD and to learn more about the confinement mechanism. They are easily accessible analytically and at the same time very accurate Monte Carlo predictions can be obtained .
Despite the availability of a wealth of information on fundamental potentials, only few lattice investigations of forces between sources in higher representations of $`SU(N)`$ gauge groups exist. Most of these studies have been performed in $`SU(2)`$ gauge theory in three and four space-time dimensions. Zero temperature results for four dimensional $`SU(3)`$ can be found in Refs. while determinations of Polyakov line correlators in non-fundamental representation have been performed at finite temperature by Bernard for $`SU(2)`$ and in Refs. for $`SU(3)`$ gauge theory.
In our study we shall see that the so-called Casimir scaling hypothesis is rather accurately represented by the lattice data while models predicting a different behaviour are definitely ruled out. Casimir scaling means that potentials between sources in different representations are proportional to each other with their ratios given by the respective ratios of the eigenvalues of the corresponding quadratic Casimir operators, which is exact in the case of two dimensional Yang-Mills theories. Our result is of particular interest with respect to recent discussions of the confinement scenario .
At distances $`r>r_c1.2`$ fm non-fundamental sources will be screened and โstring breakingโ effects will be encountered that are incompatible with Casimir scaling. In the present study we restrict ourselves to distances smaller than the string breaking scale $`r_c`$.
This article is organised as follows: in Sec. II the lattice methods that we apply and our notations are introduced. A determination of the renormalised anisotropies and lattice spacings is presented in Sec. III. The potentials are then determined in Sec. IV before we conclude with a brief discussion.
## II Notations and Methods
We denote the energy of colour sources, separated by a distance $`r`$, in a representation $`D=\mathrm{๐},\mathrm{๐},\mathrm{๐},\mathrm{๐๐},\mathrm{}`$ of the $`SU(3)`$ gauge group by $`V_D(r,\mu )`$, where $`\mu `$ denotes some cut-off scale on the gluon momenta, for instance an inverse lattice spacing, $`\mu =\pi /a`$. We shall also use the subscript โ$`F`$โ to label the fundamental ($`\mathrm{๐}`$) representation (or we may just omit the subscript in this case).
The static potential,
$$V_D(r,\mu )=V_D(r)+V_{D,\text{self}}(\mu ),$$
(1)
can be factorised into an interaction part, $`V_D(r)`$, and a self energy contribution, $`V_{D,\text{self}}(\mu )`$ that will diverge like $`\mu /\mathrm{ln}\mu `$ as $`\mu \mathrm{}`$ while $`V_D(r)`$ will assume universal values.
A (dimensionless) lattice potential $`\widehat{V}_D(๐,a)`$ will resemble the corresponding continuum potential up to lattice artefacts,
$`V_D(Ra)`$ $`=`$ $`a^1\left[\widehat{V}_D(๐,a)\widehat{V}_{D,\text{self}}(a)\right]`$ (2)
$`\times `$ $`\left[1+f_D(Ra,\widehat{๐})a^\nu \right],`$ (3)
where $`\nu `$ is a positive integer number that will in general depend on the lattice action employed. We are concerned with Wilson-type gluonic actions . In this case, $`\nu =2`$. Note that the coefficient function $`f`$ only depends on the combination $`r=Ra`$ and on the direction of $`๐`$ but not on $`R`$ itself. This guarantees that lattice artefacts are reduced as $`ra`$ is increased.
The static potentials are obtained from fits to smeared Wilson loops for $`TT_{\mathrm{min}}`$ where $`T_{\mathrm{min}}`$ depends on $`๐`$, the statistical errors of the Wilson loops and the smearing algorithm employed. We define a Wilson loop in representation $`D`$,
$$W_D(๐,T)=\text{Tr}\left(\underset{(n,\mu )\delta C(๐,T)}{}U_{D,n,\mu }\right),$$
(4)
where $`\delta C(๐,T)`$ denotes the oriented boundary of a (generalised) rectangle with spatial extent $`๐`$ and a temporal separation of $`T`$ lattice spacings. $`(n,\mu )`$ denotes an oriented link connecting the site $`n`$ with $`n+\widehat{\mu }`$, $`n`$ is an integer four-vector that labels a lattice site and $`\widehat{\mu }`$ is a unit vector pointing into a direction, $`\mu \{1,2,3,4\}`$. โTr โ is the normalised trace, $`\text{Tr}_D\mathrm{๐}_D=\frac{1}{N_D}\text{tr}\mathbf{\hspace{0.17em}1}_D=1`$, $`N_D`$ is the dimension of the representation $`D`$ and,
$$U_{D,n,\mu }=๐ซ\left\{\mathrm{exp}\left[i_{an}^{a(n+\widehat{\mu })}๐x_\mu A_\mu ^a(x)T_a^D\right]\right\},$$
(5)
denotes a link variable in representation $`D`$, where $`T_a^D`$ is a generator of the gauge group. Our conventions are $`[T_a^D,T_b^D]=if_{abc}T_c^D`$, where $`f_{abc}`$ are totally antisymmetric real structure constants. The normalisation is such that $`\text{tr}T_a^FT_b^F=\delta _{ab}/2`$. Now:
$$W_D(๐,T)=c_D(๐,a)\mathrm{exp}\left[\widehat{V}_D(๐,a)T\right](T\mathrm{}).$$
(6)
The use of smeared Wilson loops turns out to be more suitable for numerical simulations than implementing the definition of Eq. (4); the spatial pieces of the Wilson loop are replaced by linear combinations of various paths that models the ground state wave function. As a result the overlap of the creation operator with this ground state, $`c_D`$, is enhanced and the $`T\mathrm{}`$ limit can effectively be realised at moderate $`T`$ values.
In numerical simulations one observes that the statistical error $`\mathrm{\Delta }W(T)`$ of the expectation value of a smeared Wilson loop $`W(T)`$ only weakly varies with $`T`$ . From Eqs. (1) and (6) we therefore obtain the relation,
$$\frac{\mathrm{\Delta }W(T)}{W(T)}\mathrm{exp}[\widehat{V}(a)T]\mathrm{exp}\left[\widehat{V}_{\text{self}}(a)T\right](T\mathrm{}),$$
(7)
for the relative errors (which are directly proportional to the statistical uncertainty of the potential values). In tree level perturbation theory one finds,
$$\widehat{V}_{D,\text{self}}(a)=cC_Dg^2(a),c=0.252731\mathrm{};$$
(8)
the self energy is proportional to the eigenvalue of the quadratic Casimir operator $`C_D=\text{Tr}_DT_a^DT_a^D`$ of the representation. This means that statistical errors will increase significantly as we investigate higher representations of the sources with bigger Casimir charges.
This self energy related problem motivates us to introduce an anisotropy parameter $`\xi =a_\sigma /a_\tau 4`$ between spatial lattice resolution $`a_\sigma `$ and temporal lattice constant $`a_\tau `$. This results in a reduction of the self energy, $`\widehat{V}_{D,\text{self}}=cC_D\xi ^1g^2`$, and therefore of the relative errors of smeared Wilson loops. However, at the end of the day we wish to measure distance and potential in the same units $`a_\sigma `$. This means that we do not gain anything from the factor $`\xi `$ within the above expression. Nevertheless, a $`\xi >1`$ still results in a reduced effective $`g^2`$ at fixed $`a_\sigma `$. Of course, equally well we could just have increased the size of our statistical ensemble by a factor four and worked at $`\xi =1`$.
Our main motivation for introducing an anisotropy is the possibility of reducing lattice artefacts. These are most prominent at small and at large distances: as long as $`r`$ is not much larger than $`a_\sigma `$ the cubic lattice structure is clearly visible. In lowest order perturbation theory these violations of rotational symmetry only depend on $`๐`$ and $`a_\sigma `$ while the order $`g^4`$ coefficients exhibit a weak dependence on $`\xi `$ too . While we cannot hope to significantly reduce these small distance effects without decreasing $`a_\sigma `$, it is clear that on a lattice with temporal resolution $`a_\tau `$ one cannot reliably resolve masses $`ma_\tau ^1`$. However, at $`ra_\sigma `$ and in particular for representations $`D`$ with large Casimir charges situations, $`V_D(r)a_\sigma 1`$, are easily encountered, unless $`\xi 1`$.
Introducing an anisotropy also means that within any physical $`t`$ window we have more data points at our disposal. While this might help to gain more confidence in identifying effective mass plateaus we find that the additional data points are highly correlated and add little extra information, at least when one is only interested in the mass of the ground state within a given channel.
Our action reads,
$$S=\beta \underset{n}{}\left[\frac{1}{\xi _0}\underset{i>j}{}\text{Tr}U_{n,ij}+\xi _0\underset{i}{}\text{Tr}U_{n,i4}\right],$$
(9)
where $`\beta =2N/[g^2(a_\sigma ,\xi )]`$ is defined through the lattice coupling $`g^2`$ and $`i,j\{1,2,3\}`$. $`U_{n,\mu \nu }=U_{n,\mu }U_{n+\widehat{\mu },\nu }U_{n+\widehat{\nu },\mu }^{}U_{n,\nu }^{}`$ denotes the product of four link variables around an elementary square, the plaquette. With the above anisotropic Wilson action, the leading order lattice artefacts are proportional to $`a_\sigma ^2`$ and $`a_\tau a_\sigma =a_\sigma ^2\xi ^1`$. This means that along a trajectory of constant $`\xi `$ the continuum limit will be approached quadratically in $`a_\sigma `$. The relationship between the bare anisotropy $`\xi _0`$ appearing in the action Eq. (9) and the renormalised anisotropy $`\xi `$ is known in one loop perturbation theory : $`\xi =\xi _0[1+c_1(\xi _0)g^2+\mathrm{}]`$. In Sec. III we will discuss our non-perturbative evaluation of $`\xi `$. Of course the function $`\xi (g^2,\xi _0)`$ is not unique and different non-perturbative definitions might differ from each other by terms of order $`a_\sigma `$.
Perturbation theory yields the relation between potentials in different representations $`D`$,
$$V_D(r,\mu )=d_DV_F(r,\mu ),$$
(10)
where $`d_D=C_D/C_F`$. Table I contains all representations $`D`$, the corresponding weights $`(p,q)`$ and the ratios of Casimir factors, $`d_D`$, for $`p+q4`$. In $`SU(3)`$ we have $`C_F=4/3`$, and $`z=\mathrm{exp}(2\pi i/3)`$ denotes a third root of 1. Eq. (10) is known to hold to (at least) one loop (order $`g^4`$) perturbation theory at finite lattice spacing and two loops (order $`g^6`$) in the continuum limit of four dimensional Yang-Mills theories. The main purpose of this article is to investigate non-perturbatively to what extent the โCasimir scalingโ relation Eq. (10) is violated.
In what follows $`U`$ denotes a group element in the fundamental representation of $`SU(3)`$, for instance the product of link variables around a closed contour. The traces of $`U_D`$ in various representations, $`V_D=\text{tr}U_D`$, can easily be expressed in terms of traces of powers of $`U`$,
$`V_3`$ $`=`$ $`\text{tr}U,`$ (11)
$`V_8`$ $`=`$ $`\left(|V_3|^21\right),`$ (12)
$`V_6`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[(\text{tr}U)^2+\text{tr}U^2\right],`$ (13)
$`V_{15a}`$ $`=`$ $`\text{tr}U^{}V_6\text{tr}U,`$ (14)
$`V_{10}`$ $`=`$ $`{\displaystyle \frac{1}{6}}\left[(\text{tr}U)^3+3\text{tr}U\text{tr}U^2+2\text{tr}U^3\right],`$ (15)
$`V_{24}`$ $`=`$ $`\text{tr}U^{}V_{10}V_6,`$ (16)
$`V_{27}`$ $`=`$ $`|V_6|^2|V_3|^2,`$ (17)
$`V_{15s}`$ $`=`$ $`{\displaystyle \frac{1}{24}}[(\text{tr}U)^4+6(\text{tr}U)^2\text{tr}U^2+3(\text{tr}U^2)^2`$ (18)
$`+`$ $`8\text{tr}U\text{tr}U^3+6\text{tr}U^4].`$ (19)
Note that $`\text{Tr}_DU_D=\frac{1}{N_D}\text{tr}U_D=V_D/N_D`$. Hence, the normalisation of $`V_D`$ differs by a factor $`N_D`$ from that of the Wilson loop $`W_D`$ of Eq. (4). Under the replacement, $`UzU`$, $`W_D`$ transforms like, $`W_Dz^{pq}W_D`$. Representations with $`z^{pq}=1`$ have zero triality.
## III Determining anisotropy and lattice spacing
We simulate $`SU(3)`$ gauge theory at the parameter values $`(\beta ,\xi _0)=(5.8,3.10),(6.0,3.20)`$ and $`(6.2,3.25)`$. From exploratory simulations with limited statistics and the published data of Refs. we expect to find renormalised anisotropies $`\xi 4`$ at these combinations. At the above parameter values volumes of $`L_\sigma ^3\times L_\tau =8^3\times 48`$, $`12^3\times 72`$ and $`16^3\times 96`$ lattice sites have been realised, respectively. These volumes were chosen to keep the lattice extent about constant in physical units. In addition a volume of $`12^3\times 48`$ lattice sites has been simulated at $`\beta =5.8`$ to investigate possible finite size effects.
The gauge configurations have been obtained by randomly mixing Cabibbo-Marinari style Fabricius-Haan heatbath and Creutz overrelaxation sweeps, where we cycled over the three diagonal $`SU(2)`$ subgroups. The probability of a heatbath sweep was set to be $`1/5`$. During each sweep the sites were visited subsequently for each of the four space-time directions of the links in sequential order. Measurements were taken after 2000 initial heatbath sweeps and the gauge configurations are separated from each other by 200 sweeps. In doing so, we did not find any signs of autocorrelation or thermalization effects for any of the investigated observables at any of the simulated parameter sets. In the case of $`\beta =6.0`$ one set of configurations was generated on a Sparc station after a cold start while another set of configurations was generated on a Cray J90, starting from a hot, random configuration. No statistically significant deviations between these two data sets were found either. We display our simulation parameters in Table II. $`n_{\text{conf}}`$ denotes the number of statistically independent configurations analysed in each case while $`r_00.5`$ fm is the Sommer scale parameter , implicitly defined through the static potential,
$$\frac{dV(r)}{dr}|_{r=r_0}=1.65.$$
(20)
We label quantities associated with the fine grained direction by an index $`\tau `$ while $`\sigma `$ refers to coarse grained directions. On an anisotropic lattice various ways of associating the sides of smeared Wilson loops with these directions exist: $`W_{\tau \sigma }(r/a_\tau ,t/a_\sigma )`$, $`W_{\sigma \sigma }(๐ซ/a_\sigma ,t/a_\sigma )`$ and $`W_{\sigma \tau }(๐ซ/a_\sigma ,t/a_\tau )`$. In the case of $`W_{\tau \sigma }`$ as well as for $`W_{\sigma \sigma }`$ the time coordinate points into a $`\sigma `$ direction. The spatial coordinate is identified with the $`\tau `$ direction in the first case and a $`\sigma `$ direction in the latter case. While the spatial connections within $`W_{\tau \sigma }`$ are parallel to the $`\tau `$ axis, in the case of $`W_{\sigma \sigma }`$ we realise planar off-axis configurations $`๐(1,1,0)`$ and $`๐(2,1,0)`$, in addition to on-axis separations<sup>*</sup><sup>*</sup>* We ignore the possibility of mixing $`\tau `$ and $`\sigma `$ coordinates within the โspatialโ separation.. Finally, within $`W_{\sigma \tau }`$ the time coordinate is taken along the fine grained dimension and the spatial coordinate coarse grained. We determine $`W_{\sigma \tau }`$ for the standard separations , $`๐(1,0,0),`$ $`(1,1,0),`$ $`(2,1,0),`$ $`(1,1,1),`$ $`(2,1,1),`$ $`(2,2,1)`$.
In the spatially isotropic situation ($`W_{\sigma \tau }`$) we iteratively construct fat links in the standard way by replacing a given link by the sum of itself and the neighbouring four spatial staples with some weight parameter, $`\alpha 1`$,
$`U_{n,i}`$ $``$ $`P_{SU(3)}\left(\alpha U_{n,i}+{\displaystyle \underset{ji}{}}F_{n,j}\right),`$ (21)
$`F_{n,j}`$ $`=`$ $`U_{n,j}U_{n+\widehat{ศท},i}U_{n+\widehat{ฤฑ},j}^{}+U_{n\widehat{ฤฑ},i}^{}U_{n\widehat{ฤฑ},i}U_{n\widehat{ฤฑ}+\widehat{ศท},j}.`$ (22)
$`P_{SU(3)}`$ denotes a projection operator, back onto the $`SU(3)`$ manifold. We employ the definition , $`U=P_{SU(3)}(A)SU(3)`$, $`\text{Re}\text{Tr}UA^{}=\mathrm{max}`$ and iterate Eq. (21) 26 times with $`\alpha =2.3`$.
We use a somewhat different novel smearing algorithm in the case of $`W_{\tau \sigma }`$ and $`W_{\sigma \sigma }`$ where the spatial volume is anisotropic with one fine and two coarse directions: when one only considers links parallel to the one being replaced, Eq. (21) resembles a two dimensional diffusion process: $`UP_{SU(3)}[(\alpha +4+_2^2)U]`$. We are interested to maintain an isotropic propagation of the link fields when an anisotropy parameter $`\xi >1`$ is introduced. Following the above diffusion model this is achieved by replacing Eq. (21) with,
$`U_{n,i}`$ $``$ $`P_{SU(3)}\left(\alpha U_{n,i}+F_{n,j}+\xi ^2F_{n,\tau }\right),`$ (23)
$`U_{n,\tau }`$ $``$ $`P_{SU(3)}\left[(\alpha +2\xi ^22)U_{n,\tau }+{\displaystyle \underset{i}{}}F_{n,i}\right],`$ (24)
where $`i,j\{1,2\},ji`$. We perform 22 iterations of Eqs. (23) โ (24) with $`\alpha =3.7`$. Indeed, in doing so we find very similar overlaps with the physical ground state, $`c_{\sigma \sigma }(๐)c_{\tau \sigma }(๐)`$, where $`c=c_F[0,1]`$ is defined in Eq. (6) and $`W_{\tau \sigma }(๐,T=0)=W_{\sigma \sigma }(๐,T=0)=1`$. This is illustrated in the comparison of data obtained at $`\beta =6.2`$, $`\xi _0=3.25`$ of Fig. 1. We have not been able, however, to sustain the high overlaps achieved for $`W_{\sigma \tau }`$ (triangles) for $`W_{\tau \sigma }`$ or $`W_{\sigma \sigma }`$. The situation at the other $`(\beta ,\xi _0)`$ combinations is similar.
From the asymptotic behaviour of the different Wilson loops at large temporal separations $`t`$, three lattice potentials can be determined:
$`a_\sigma ^1\widehat{V}_{\tau \sigma }(r/a_\tau )`$ $`=`$ $`\underset{t\mathrm{}}{lim}{\displaystyle \frac{d}{dt}}\mathrm{ln}W_{\tau \sigma }(r/a_\tau ,t/a_\sigma ),`$ (25)
$`a_\sigma ^1\widehat{V}_{\sigma \sigma }(๐ซ/a_\sigma )`$ $`=`$ $`\underset{t\mathrm{}}{lim}{\displaystyle \frac{d}{dt}}\mathrm{ln}W_{\sigma \sigma }(๐ซ/a_\sigma ,t/a_\sigma ),`$ (26)
$`a_\tau ^1\widehat{V}_{\sigma \tau }(๐ซ/a_\sigma )`$ $`=`$ $`\underset{t\mathrm{}}{lim}{\displaystyle \frac{d}{dt}}\mathrm{ln}W_{\sigma \tau }(๐ซ/a_\sigma ,t/a_\tau ).`$ (27)
Note that while $`\widehat{V}_{\tau \sigma }`$ and $`\widehat{V}_{\sigma \sigma }`$ are given in units of $`a_\sigma `$, $`\widehat{V}_{\sigma \tau }`$ is measured in units of $`a_\tau `$. These potentials are related to each other:
$`\widehat{V}_{\sigma \sigma }(R,a_\sigma )`$ $`=`$ $`\widehat{V}_{\tau \sigma }(\xi R,a_\sigma )[1+๐ช(a_\sigma )^2]`$ (28)
$`=`$ $`\left[\xi \widehat{V}_{\sigma \tau }(R,a_\sigma )+\mathrm{\Delta }\widehat{V}_{\text{self}}(a_\sigma )\right][1+๐ช(a_\sigma )^2],`$ (29)
where $`\mathrm{\Delta }\widehat{V}_{\text{self}}=\widehat{V}_{\sigma \sigma ,\text{self}}\xi \widehat{V}_{\sigma \tau ,\text{self}}`$. While $`\widehat{V}_{\sigma \sigma }`$ and $`\widehat{V}_{\tau \sigma }`$ are equal at a given physical distance (up to lattice artefacts), in the case of $`\widehat{V}_{\sigma \tau }`$ a shift by an additive constant is expected since the self energies differ:
$$\mathrm{\Delta }\widehat{V}_{\text{self}}(a_\sigma )=0.08214\mathrm{}g^2+\mathrm{}.$$
(30)
The numerical value has been obtained in lowest order perturbation theory for $`\xi =\xi _0=4`$.
Following Ref. we use Eq. (28) to determine the renormalised anisotropiesNote that unlike Ref. our analysis is based on asymptotic $`T\mathrm{}`$ results rather than on pre-asymptotic finite $`T`$ approximants to the potential.. Eq. (29) can then be used as an independent consistency check (modulo lattice artefacts). In order to guarantee a consistent definition of $`\xi `$ we can either consider the limit $`R\mathrm{}`$ or demand the matching to be performed at the same distance in terms of a measured correlation length. We follow the latter strategy and impose,
$$\widehat{V}_{\sigma \sigma }(R_m^L,a_\sigma )=\widehat{V}_{\tau \sigma }(\xi R_m^L,a_\sigma ),$$
(31)
at $`R_m^LR_m=(2/3)r_0/a_\sigma `$ where $`r_00.5`$ fm is the Sommer scale of Eq. (20). We restrict ourselves to on-axis separations and take $`R_m^L=2,3,4`$ for $`\beta =5.8,6.0`$ and $`6.2`$, respectively. This choice is justified by our subsequent analysis where we find $`R_m=2.05(2)`$ and $`R_m=2.03(2)`$ on the $`8^3`$ and $`12^3`$ $`\beta =5.8`$ lattices and $`R_m=2.96(4)`$ and $`R_m=4.07(4)`$ at $`\beta =6.0`$ and $`\beta =6.2`$, respectively. The renormalised anisotropy $`\xi `$ is then obtained from Eq. (31) by interpolating the potential $`V_{\tau \sigma }`$ according to three parameter fits,
$$\widehat{V}_{\tau \sigma }(R)=\widehat{V}_{0,\tau \sigma }+K_{\tau \sigma }R\frac{e_{\tau \sigma }}{R}.$$
(32)
The errors are obtained via the bootstrap procedure. The resulting anisotropies and fit ranges employed, $`R[R_{\mathrm{min}},L_\tau /2]`$, as well as the fit parameters in units of $`a_\sigma `$ are displayed in Table III. $`T_{\mathrm{min}}=t_{\mathrm{min}}/a_\sigma `$ denotes the โtemporalโ separation from which onwards effective mass data \[Eqs. (25) and (26)\] saturated into plateaus.
The renormalised anisotropies $`\xi `$ are also included in the last column ofNote that in this table we have averaged the results obtained on the $`8^3`$ and $`12^3`$ lattices at $`\beta =5.8`$ that agree with each other within errors. Table IV. In the third column of this table the one loop results are displayed while in the second last column mean field estimates are shown,
$$\xi _{ir}=\xi _0\sqrt{\frac{U_{\sigma \tau }}{U_{\sigma \sigma }}},\beta _{ir}=\beta \sqrt{U_{\sigma \tau }U_{\sigma \sigma }}.$$
(33)
The temporal and spatial average plaquettes ($`U_{\sigma \tau }`$ and $`U_{\sigma \sigma }`$) are also included in the table. While the renormalised anisotropies are underestimated by the one loop results by about 10 % they are overestimated by the mean field values by almost the same amount. Finally, in Table V we compile the bare anisotropies $`\xi _{0,4}=4\xi _0/\xi `$ at which we should have simulated in order to achieve $`\xi =4`$. In doing so we assume that our statistical uncertainties on $`\xi `$ of order 1 % will dominate over variations of the ratios $`\xi _0/\xi `$ under a change of $`\xi _0`$ by less than 2 %.
After determining the anisotropies, the potential $`V_{\sigma \tau }`$ is fitted to the parametrisation Eq. (32) for $`rr_m^Lr_m=2r_0/3`$. The results are compiled in Table VI. The fit parameters $`e`$ and $`K`$ agree within errors with those determined from the data on $`V_{\tau \sigma }`$ of Table III while the $`V_0`$ values tend to be somewhat smaller, in agreement with the expectation of Eq. (30). Note that the parametrisation is thought to be effective only and that the fit ranges employed for the two potentials differ from each other. From the fit parameters, values $`r_0/a_\sigma =\sqrt{(1.65\xi ^1e)/K}`$ can be extracted. These are displayed in Table II, along with the linear spatial lattice extent. Compared to the isotropic case, $`\xi _0=1`$, where $`r_0/a=3.64(5),5.33(3)`$ and $`7.29(4)`$ at $`\beta =5.8,6.0`$ and $`6.2`$, respectively, $`a_\sigma `$ is somewhat increased while the temporal lattice spacing $`a_\tau `$ is reduced. The ratios $`r_0a_\sigma ^{3/4}a_\tau ^{1/4}=4.36(4),`$ $`6.31(10)`$ and $`8.61(9)`$ exhibit that at $`\xi =4`$ the geometrically averaged lattice spacings are about 15 % smaller than their isotropic counterparts, obtained at the same $`\beta `$ values.
Assigning the phenomenological value $`0.5`$ fm to $`r_0`$ we find $`L_\sigma a_\sigma 1.3`$ fm on the small lattices and $`L_\sigma a_\sigma 2`$ fm on the $`12^3`$ lattice at $`\beta =5.8`$. This means that $`\sqrt{3}L_\sigma a_\sigma /2>1.1`$ fm in all our simulations; along the $`๐(1,1,1)`$ direction we are safe from the effect of mirror charges up to distances bigger than one fm. Beyond this distance only representations with non-zero triality are protected by the centre symmetry from direct finite size effects.
In Fig. 2 we display all three potentials in units of $`a_\sigma `$ at $`\beta =6.2`$. Note that the anisotropy has been determined by matching $`V_{\tau \sigma }`$ to $`V_{\sigma \sigma }`$ at $`r=4a_\sigma `$. In addition to the data points two curves are included that correspond to the parameter values of Table III and Table VI from fits according to Eq. (32) to $`V_{\tau \sigma }`$ for $`r>2a_\sigma `$ and to $`V_{\sigma \tau }`$ for $`r4a_\sigma `$, respectively. The matched potentials $`V_{\sigma \sigma }`$ and $`V_{\tau \sigma }`$ follow the same curve.
Up to lattice artefacts and the self energy shift $`\mathrm{\Delta }\widehat{V}_{\text{self}}`$, $`\xi \widehat{V}_{\sigma \tau }`$ and $`\widehat{V}_{\sigma \sigma }`$, that both live along coarse grained lattice directions, should also agree with each other \[Eq. (29)\]. Indeed, as is demonstrated in Fig. 3, the differences are compatible with constants of the order suggested by tree level perturbation theory for $`\xi =\xi _0=4`$, Eq. (30). Averaging the $`RR_m^L`$ data points results in the values, $`0.091(9)`$, $`0.096(18)`$ and $`0.124(11)`$ for $`\mathrm{\Delta }\widehat{V}_{\text{self}}`$ at $`\beta =5.8,6.0`$ and 6.2, respectively (solid lines with error bands). On the large lattice at $`\beta =5.8`$ we obtain the value $`0.081(9)a_\sigma ^1`$, in agreement with that above, from the smaller volume. These shifts of the self energies result in reduced relative errors of $`W_{\sigma \tau }`$ \[Eq. (7)\] (and in increased errors of $`W_{\sigma \sigma }`$), relative to the isotropic case.
## IV The potentials
The potentials in non-fundamental representations are extracted in the same way as discussed above from fits to the corresponding smeared Wilson loops. These are obtained from the fundamental ones by use of Eqs. (11) โ (19). In the case of the fundamental Wilson loops, discussed in Sec. III, temporal links have been replaced by their thermal averages in the vicinity of the surrounding staples in order to reduce statistical fluctuations (link integration ). Note that our use of Eqs. (11) โ (19) implies that we cannot thermally average fundamental links in the construction of higher representation Wilson loops anymore.
We determine the potentials from correlated exponential fits to $`W_{D,\sigma \tau }`$ data according to Eq. (6). The fit range in $`T`$ is selected separately for each distance $`๐`$ and representation $`D`$, such that $`\chi ^2/N_{DF}<1.5`$. In addition, we demand the saturation of โeffective massesโ, $`V_D(T)a_\sigma =\frac{\xi }{4}\mathrm{ln}[W_D(T)/W_D(T+4)]`$, into plateaus for $`TT_{\mathrm{min}}`$. In Table VII we display the resulting fit ranges $`T[T_{\mathrm{min}},T_{\mathrm{max}}]`$ that have been selected by means of this procedure for the example of the point $`\widehat{V}_D(๐)`$ with $`Rr_0/a_\sigma `$. In general, the interplay between statistical errors and ground state overlaps was such that $`T_{\mathrm{min}}`$ only slightly varied with $`R`$. In the case of the fundamental potential we find values $`2r_0t_{\mathrm{min}}4.5r_0`$, depending on $`R`$ and the parameter values we simulate at, while for $`D=\mathrm{๐๐}๐ฌ`$ we find, $`0.7r_0t_{\mathrm{min}}r_0`$. The corresponding estimated ground state overlaps $`c_D`$ are displayed in Table VIII. The overlaps decrease with increasing Casimir constant, lattice spacing or distance $`r`$. At $`rr_0`$ the overlaps range from $`0.62(3)`$ in the worst case to $`0.97(1)`$ in the best case which quantifies the efficiency of our smearing algorithm.
In Fig. 4 we display the resulting potentials at $`\beta =6.2`$. The curves correspond to the three parameter fit Eq. (32) to the fundamental potential, multiplied by the respective ratio of Casimir factors, $`d_D`$ of Table I. It is clear that the Casimir scaling hypothesis Eq. (10) works quite well on our $`\beta =6.2`$ data for the investigated distances. In Figs. 57 we display the ratios $`\widehat{V}_D(R,a_\sigma )/\widehat{V}_F(R,a_\sigma )`$ for our three lattice spacings. We did not attempt to subtract the self energy contributions \[cf. Eqs. (1) โ (2)\] in this comparison. At $`\beta =5.8`$ and $`\beta =6.0`$ we find the data to lie significantly below the corresponding Casimir ratios (horizontal lines). However, the deviations decrease rapidly as the lattice spacing is reduced.
Prior to a continuum limit extrapolation of the ratios we shall investigate finite size effects by comparing results obtained on the 1.3 fm lattice with results from the 2 fm lattice at $`\beta =5.8`$. For the fundamental potential we already know from previous studies that for spatial extents, $`L_\sigma a_\sigma >2r_0`$, such effects are well below the 2 % level . The situation is less clear for potentials in higher representations. In principle, the flux tube between the sources could widen when the energy per unit length is increased and, therefore, higher representation potentials might be more susceptible to finite size effects.
In Fig. 8 we compare the fundamental, octet and sextet potentials obtained on the $`12^3`$ lattices at $`\beta =5.8`$ (full symbols) with those obtained on the $`8^3`$ lattices (open symbols). Up to distances well beyond $`2r_06a_\sigma `$ no statistically significant deviations are seen. In Fig. 9 we show the relative deviations, $`V_D^{L_\sigma =12}(๐ซ)/V_D^{L_\sigma =8}(๐ซ)1`$, between the potentials determined on the larger lattices and those measured on the smaller lattices for all the representations that we have investigated. Again, no systematic or statistically significant differences are detected. Up to $`r=4a_\sigma 1.4r_0`$ this holds true on the 1 % level for the fundamental potential and on the 3โ5 % level for higher representation potentials. Beyond this distance the statistical errors start to explode. The same comparison has been performed for ratios of potentials. In this case the relative errors are slightly reduced due to correlation effects. However, no statistically significant tendencies were observed either. The relative statistical errors on the two lattices are of about the same size and comparable to those of the $`\beta =6.0`$ and $`\beta =6.2`$ simulations. Thus, we do not expect finite size effects to exceed the statistical errors on any of the simulated lattice volumes.
We now attempt a continuum limit extrapolation of our data. We remark that in the limit $`a_\sigma 0`$ \[Eqs. (1) โ (2)\] the Casimir scaling of the diverging self energies $`\widehat{V}_{D,\text{self}}(a_\sigma )`$ automatically implies Casimir scaling of $`\widehat{V}_D(๐,a_\sigma )`$. However, with $`V_{D,\text{self}}`$ being a purely ultra violet quantity, this sort of Casimir scaling has little to do with non-perturbative aspects of the theory. As can be seen from Fig. 4, where the potentials vary by more than a factor two with the distance, even at our finest lattice resolution $`V_{D,\text{self}}`$ have not yet become the dominant contributions to the lattice potentials. In order to avoid the trivial Casimir scaling described above we will only study ratios of physical interaction energies,
$$R_D(r)=\frac{V_D(r)}{V_F(r)}=\frac{\widehat{V}_D(๐,a_\sigma )\widehat{V}_{D,\text{self}}(a_\sigma )}{\widehat{V}_F(๐,a_\sigma )\widehat{V}_{F,\text{self}}(a_\sigma )}\left[1+๐ช(a_\sigma ^2)\right],$$
(34)
where $`r=Ra_\sigma `$.
We estimate the self energies in leading order perturbation theory, Eq. (8). Lattice perturbation theory is notorious for its bad convergence behaviour . However, in our anisotropic case the effective expansion parameter $`g^2\xi _0^1`$ is much smaller than in standard applications of lattice perturbation theory. In Fig. 3 we have indeed seen that leading order perturbation theory predictions on the difference of two self energies agree within 30 % with numerical data. We estimate the self energies in two different ways: (a) we use the bare lattice coupling and the renormalised anisotropy $`\xi `$, (b) we mean field (โtadpoleโ) improve both, coupling and anisotropy \[Eq. (33)\], $`\xi _{0,ir}^1g_{ir}^2=\xi _0^1g^2U_{\sigma \tau }^1`$. The results from the two methods, shown in Table IX, differ by up to 60 % from each other. We will use the estimate (a) in our analysis but take the difference between (a) and (b) into account as a systematic error. While data at large distances are marginally affected by this uncertainty, the error bars at small distances are significantly increased.
After subtracting the (scaling violating) self energy contributions we determine the continuum extrapolated ratios $`R_D(r)`$ by means of quadratic fits, Eq. (34). We perform these extrapolations for all the distances $`r`$ that have been realised on our coarsest lattice ($`\beta =5.8`$) in units of $`r_0`$. On the finer lattices, we linearly interpolate between the two lattice points that are closest to each given distance $`r`$, prior to the quadratic continuum limit fit. We find the data to be compatible with the quadratic ansatz and the resulting ratios are shown in Fig. 10. The numerical values are displayed in Tables XXI. No statistically significant violations of Casimir scaling are found. Our accuracy is somewhat limited at short distances, due to the perturbative estimation of the self energies. The slope of the extrapolation in $`a_\sigma ^2`$ increases with the distance $`r`$ as well as with the Casimir charge $`C_D`$; large masses $`\widehat{V}_D(R,a_\sigma )>a_\tau ^1`$ are more affected by lattice artefacts than small masses. This observation also explains why the deviations from Casimir scaling at $`\beta =5.8`$ and $`\beta =6.0`$ increase with the distance (Figs. 56).
## V Discussion
We have confirmed that violations of the Casimir scaling hypothesis,
$$\frac{V_{D_1}(r)}{V_{D_2}(r)}=\frac{C_{D_1}}{C_{D_2}},$$
(35)
are below the 5 % level for distances $`r<2r_01`$ fm in the continuum limit of four dimensional $`SU(3)`$ gauge theory for all representations with $`C_D7`$. This finding rules out many models of non-perturbative QCD and imposes serious restrictions onto others. For instance in a bag model calculation scaling of string tensions with the square root of the respective Casimir ratio has been obtained and instanton liquid calculations result in ratios between potentials in different representations that are smaller than the Casimir ratios too .
Another possibility would have been scaling proportional to the number of fundamental flux tubes embedded into the higher representation vortex \[$`p+q`$ in $`SU(3)`$, which happens to coincide with Casimir scaling in the large $`N`$ limit of $`SU(N)`$. This picture is supported by the finding that the $`SU(N)`$ vacuum seems to act like a type I superconductor , i.e. flux tubes repel each other. However, this scenario is also excluded by the present study. Furthermore, serious restrictions onto most of the remaining models are imposed (see for instance Ref. ). It is particularly disappointing that neither centre vortex models nor the dual superconductor scenario or string models seem to offer any explanation why the numerical data so closely resemble the Casimir ratios. Certainly, it is worthwhile to dedicate more theoretical effort to this fundamental phenomenon.
We have not discussed โstring breakingโ so far. While the fundamental potential in pure gauge theories linearly rises ad infinitum, the adjoint potential will be screened by gluons and, at sufficiently large distances, decay into two disjoint gluelumps . This string breaking has indeed been confirmed in numerical studies . Therefore, strictly speaking, the adjoint string tension is zero. In fact, all charges in higher than the fundamental representation will be at least partially screened by the background gluons. For instance, $`\mathrm{๐}\mathrm{๐}=\mathrm{๐๐}\mathrm{๐๐}๐^{}\mathrm{๐}\mathrm{๐}^{}`$: in interacting with the glue, the sextet potential obtains a fundamental component. A simple rule, related to the centre of the group, is reflected in Eqs. (11) โ (19): wherever $`z^{pq}=1`$ (zero triality), the source will be reduced into a singlet component at large distances while, wherever $`z^{pq}=z`$ (or $`z^{}`$), it will be screened, up to a residual (anti-)triplet component, i.e. one can easily read off the asymptotic string tension (either zero or the fundamental string tension) from the third column of Table I, rather than having to multiply and reduce representations. As a result, the self-adjoint representations, $`\mathrm{๐}`$ and $`\mathrm{๐๐}`$, as well as the representation, $`\mathrm{๐๐}`$, will be completely screened while in all other representations with $`p+q4`$ a residual fundamental component survives. The same argument, applied to $`SU(2)`$, results in the prediction that all odd-dimensional (bosonic) representations are completely screened while all even-dimensional (fermionic) representations will tend towards the fundamental string tension at large distances.
One expects this sort of string breaking and flattening of the potential to occur at distances larger than about 2.4 $`r_0`$ . Obviously, once the string is broken Casimir scaling is violated. It is certainly interesting to investigate what happens around the string breaking distance. However, this requires lattice volumes exceeding those used in the present study as well as additional operators that are designed for an optimal overlap with the respective broken string states .
###### Acknowledgements.
This work was supported by DFG grants Ba 1564/3-1, 1564/3-2 and 1564/3-3 as well as EU grant HPMF-CT-1999-00353. The simulations were performed on the Cray J90 system of the ZAM at Forschungszentrum Jรผlich as well as on workstations of the John von Neumann Institut fรผr Computing. We thank the support teams of these institutions for their help.
|
warning/0006/hep-ph0006278.html
|
ar5iv
|
text
|
# Three flavour neutrino oscillations in models with large extra dimensions
## I Introduction
Particle physics models where there are large hidden space dimensions beyond the three familiar ones have been the focus of intense activity during the past two years. Beyond the simple reason that such extra dimensions are predicted by string theories, a major point of interest in these models is that often these large extra dimensions come with a TeV scale for the strings which leads to a plethora of new observable phenomena in collider as well in other arenas of particle physics and cosmology.
In order for these models to provide a satisfactory description of low energy particle physics, they must handle some obvious problems that come with the existence of a fundamental scale in the multi-TeV range. One such problem has to do with understanding the small neutrino masses in a natural manner. The conventional seesaw explanation which is believed to provide the most satisfactory way to understand this, requires that the new physics scale (or the scale of $`SU(2)_R\times U(1)_{BL}`$) be around $`10^{12}`$ GeV or higher. Clearly, low string scale theories do not have any fundamental scale of that type. Moreover, the low value of the string scale leads to enhanced (and unacceptable) contributions to neutrino masses from higher dimensional operators. While there are suggestions involving thick branes with point splitting to remedy similar problems that arise from $`\mathrm{\Delta }B0`$ operators, they donโt work for neutrino mass operators. Therefore, a necessary ingredient to understand small neutrino masses in the low string scale models is to assume that theory have a $`BL`$ symmetry. This will forbid higher dimensional operators $`LHLH/M^{}`$ ($`M^{}`$ is the string scale), which are the source of the problem for neutrino masses.
Depending on whether $`BL`$ is a global or local symmetry, one can have two ways to solve the neutrino mass problem in models with large extra dimensionsIt is generally believed that string theories do not have any global symmetries, which would seem to imply that $`BL`$ must be a local symmetry. In our discussion, however, we will consider that $`BL`$ is a global symmetry of the theory, as in Ref. and see where it leads us phenomenologically.. In the former case, discussed in , one has to introduce singlet bulk neutrinos which then lead to small Dirac masses for them. On the other hand, if $`BL`$ symmetry is chosen to be a local symmetry, anomaly cancellation requires that, right handed neutrinos be present in the brane as in the models discussed in Ref.. Since local $`BL`$ must be broken to avoid massless gauge bosons, one again has to deal with the induced operators of the same type as above with $`M^{}=M_{BL}`$. It was shown in Ref. that to get neutrino masses in the desired eV range, one must have string scale $`M^{}M_{BL}10^9`$ GeV range or higher. As a result, a class of experimentally accessible phenomenological predictions are lost, although long range gravity tests are still possible. These models are similar to the ones discussed in .
In the context of models that have global $`U(1)_{BL}`$ symmetry, one can maintain the TeV scale for the strings and still get small neutrino masses by introducing isosinglet neutrinos in the bulk as has already been discussed in Ref.. These models are interesting because a very minimal set of particles beyond the standard model are sufficient to get small neutrino masses. Key reason for this result is the relation between the fundamental scale, $`M^{}`$, the radius of the extra dimensions, $`R`$, and effective Planck scale, $`M_P\mathrm{}`$,
$$M^{2+n}R^n=M_P\mathrm{}^2,$$
(1)
Since the singlet neutrino is a bulk field, the effective couplings of its Fourier modes to the standard model fields are naturally suppressed by the ratio $`M^{}/M_P\mathrm{}`$, which for a TeV $`M^{}`$ produces the right order of magnitude for neutrino masses.
An analysis of the implications of the mixing profile in these models for solar neutrino deficit was discussed in . Also, implications for atmospheric neutrinos were discussed in , and some phenomenological bounds were given in . Our goal in this paper is to attempt a unified explanation of all known neutrino oscillation data i.e. solar, atmospheric as well as LSND under different assumptions for the initial input parameters for the minimal bulk neutrino scenario.
The basic reason for embarking on such an ambitious program is the encouraging feature that the masses of the lower KK modes of the bulk neutrinos are given by integral multiples of $`R^1`$ which is of order $`10^3`$ eV when $`R`$ millimeter. This is of the right order necessary to solve the solar neutrino problem via small angle MSW mechanism using $`\nu _e`$ to $`\nu _s`$ oscillation. Thus we see that there exists a natural way to understand the lightness of the sterile neutrino, a situation if realized would pose a major challenge to four dimensional theories. Once the solar neutrino problem is understood, it would appear that all the necessary ingredients are at hand to understand the atmospheric and LSND data using oscillations among the familiar neutrinos i.e. $`\nu _e\nu _\mu `$ for LSND and $`\nu _\mu \nu _\tau `$ for atmospheric.
This program has already been undertaken in special parameter domains and it has already been suspected that it does not really work. Our goal is to extend these discussions to a somewhat larger parameter domain to see if there is chance for this program to succeed and unfortunately our answer is also in the negative. Our work complements the above works and extends them. Specifically, we try to give analytical reasonings to see how the different oscillation data can (or cannot) be understood. Since we do not take recourse to a detailed numerical analysis, we cannot rule out the possibility that some small parameter domain exists where all data can be accommodated; but we consider that unlikely.
The negative conclusion of our work, combined with the works of implies with virtual certainty that in the large extra dimension framework, understanding the neutrino data does require new physics beyond the standard model in the brane or new physics in extra dimensions or both.
Our basic strategy is as follows: we start by requiring that the parameters of the theory provide an explanation for solar and atmospheric oscillations. Then we ask whether they can account for the small LSND probability of $`\nu _e`$ appearance from the $`\nu _\mu `$ beam.
This paper is organized as follows: in section II, we discuss neutrino oscillations with a single flavour, which depends on two parameters, the bulk radius and a dimensionless parameter $`\xi `$ related to the Dirac mass term in the theory. The latter defines the pattern of neutrino oscillations. This analysis sets the stage for the three flavour case, which we discuss in section III for various possible domains of the parameter space. We end the paper with a concluding section that summarizes the results and discusses their implications.
## II Oscillations with a single flavour
Let us begin our discussion by focusing on the simplest case with one generation of fermions in the brane and one bulk neutrino, to understand the general profile of the neutrino oscillations in models with large extra dimensions. We will discuss the necessary ingredients to understand the three flavour case that is explored in the next section. Obviously, the fields that could propagate in the extra dimensions are chosen to be gauge singlets. Let us denote bulk neutrino by $`\nu _B(x^\mu ,y)`$. It has a five dimensional kinetic energy term and a coupling to the brane field $`L(x^\mu )`$ given by
$$=\kappa \overline{L}H\nu _{BR}(x,y=0)+๐y\overline{\nu }_{BL}(x,y)_5\nu _{BR}(x,y)+h.c.$$
(2)
where from the five dimensional kinetic energy, we have only kept the 5th component that contributes to the mass terms of the KK modes in the brane; $`H`$ denotes the Higgs doublet, and $`\kappa =h\frac{M^{}}{M_P\mathrm{}}`$ the suppressed Yukawa coupling. It is worth pointing out that this suppression is independent of the number and radius hierarchy of the extra dimensions, provided that our bulk neutrino propagates in the whole bulk. For simplicity, we will assume that there is only one extra dimension with radius of compactification as large as a millimiter, and the rest with much smaller compactification radii. The smaller dimensions will only contribute to the relationship (1) but its KK excitations will be very heavy and decouple from neutrino spectrum. Thus, all the analysis could be done as in five dimensions.
The first term in Eq. (2) will be responsible for the neutrino mass once the Higgs field develops its vacuum. The induced Dirac mass parameter will be given by $`m=\kappa v`$, which for $`M^{}=1`$ TeV is about $`h10^5`$ eV. Obviously this value depends only linearly on the fundamental scale. Larger values for $`M^{}`$ will increase $`m`$ proportionally. After introducing the expansion of the bulk field in terms of the KK modes, the Dirac mass terms in (2) could be written as
$$(\overline{\nu }_{eL}\overline{\nu }_{BL}^{})\left(\begin{array}{cc}m& \sqrt{2}m\\ 0& _5\end{array}\right)\left(\begin{array}{c}\nu _{0B}\\ \nu _{BR}^{}\end{array}\right),$$
(3)
where our notation is as follows: $`\nu _B^{}`$ represents the KK excitations, the off diagonal term $`\sqrt{2}m`$ is actually an infinite row vector of the form $`\sqrt{2}m(1,1,\mathrm{})`$. The operator $`_5`$ stands for the diagonal and infinite KK mass matrix whose $`n`$-th entrance is given by $`n/R`$. This notation was introduced in to represent the infinite mass matrix in a compact manner.
Using this short hand notation makes it easier to calculate the exact eigenvalues and the eigenstates of this mass matrix . Simple algebra yields the characteristic equation
$$2\lambda _n=\pi \xi ^2\mathrm{cot}(\pi \lambda _n),$$
(4)
with $`\lambda _n=m_nR`$, $`\xi =\sqrt{2}mR`$, and where $`m_n`$ is the mass eigenvalue . The eigenstates, on the other hand are given symbolically by
$$\stackrel{~}{\nu }_{nL}=\frac{1}{N_n}\left[\nu _L+\frac{\sqrt{2}m_5}{m_n^2_5^2}\nu _{BL}^{}\right],$$
(5)
where the sum over the KK modes in the last term is implicit. $`N_n`$ is the normalization factor given by
$$N_n^2=\frac{1}{\xi ^2}\left(\lambda _n^2+f(\xi )\right),$$
(6)
where $`f(\xi )=\xi ^2/2+\pi ^2\xi ^4/4`$. Figures 1 and 2 depict the exact numerical results of $`\lambda _n`$ and $`N_n`$ for several choices for the value of the $`\xi `$ parameter. Using the expression (5), we can write down the weak eigenstate $`\nu _L`$ in terms of the massive modes as
$$\nu _L=\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{N_n}\stackrel{~}{\nu }_{nL}.$$
(7)
Thus, the weak eigenstate is actually a coherent superposition of an infinite number of massive modes. Therefore, even for this single flavour case, the time evolution of the mass eigenstates involves in principle all mass eigenstates and is very different from the simple oscillatory behaviour familiar from the conventional two or three neutrino case. The time dependent survival probability is given by
$`P_{surv}(L)`$ $`=`$ $`\left|\nu _L(L)|\nu _L(0)\right|^2=\left|{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{e^{i\frac{L}{2ER^2}\lambda _n^2}}{N_n^2}}\right|^2=12{\displaystyle \underset{k,n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\mathrm{sin}^2\left(\frac{L}{4ER^2}(\lambda _n^2\lambda _k^2)\right)}{N_n^2N_k^2}}.`$ (8)
It is clear that the survival probability depends strongly on the parameter $`\xi `$, reflecting the universal coupling of all the KK components of $`\nu _B`$ with $`\nu _L`$ in (2). Figures 3 and 4 show the profile of the survival probability obtained from the numerical solutions for three different values of $`\xi `$ and is clearly very different from simple familiar oscillatory behaviour. However, to better understand these results, we will follow an analytical approach in what follows.
It is simpler to consider the two limiting cases. First let us assume that $`\xi 1`$. As already known , the eigenvalues in this case are given by $`\lambda _0=mR`$, and $`\lambda _n=n`$ otherwise. Therefore, to a good approximation, we may take the mixing parameters $`N_0=\eta `$, and $`N_n=(n/\xi )\eta `$ for non zero $`n`$. Where the extra factor $`\eta =(1+\pi ^2\xi ^2/6)^{1/2}`$ is introduced to keep the proper normalization in the expansion (7). This approximation is confirmed by our figures 1 and 2. The survival probability is now given as
$$P_{surv}(L)=1\frac{4}{\eta ^4}\xi ^2\underset{n=1}{\overset{\mathrm{}}{}}\frac{\mathrm{sin}^2\left(\frac{n^2L}{4ER^2}\right)}{n^2}\frac{2}{\eta ^4}\xi ^4\underset{k,n=1}{\overset{\mathrm{}}{}}\frac{\mathrm{sin}^2\left[\frac{(n^2k^2)L}{4ER^2}\right]}{n^2k^2}.$$
(9)
It is simple to see from last expression that the probability has an oscillation length $`L_{osc}=4\pi ER^2`$. A typical profile in this case is depicted in figure 4. Also, the main contribution to the oscillation pattern comes from the lowest elements of the tower, which turns out to be the main component of $`\nu _L`$. Another way to see this result is to note that $`\xi `$ could be made small by making $`R`$ small; but this makes the KK excitation masses large so that they decouple from the light sector of the theory leaving only the right handed component of the zero mode of the bulk neutrino, $`\nu _{BR}`$ to form a Dirac mass with $`\nu _L`$ and stay light. The left-handed zero mode decouples and remains massless.
In the $`\xi 1`$ limit, the small mixing with the tower elements leads to an oscillation pattern dominated by the lightest KK mode with a mass just about $`1/R`$ and more or less simulate the familiar oscillatory one. This can happen, for instance, if $`R0.2mm`$, which gives $`1/R^210^6eV^2`$, just about what it is needed to provide an explanation to the solar neutrino problem assuming a small MSW mixing angle . To use this case to compare with solar neutrino data, one needs to include the matter effect, which has already been discussed in Ref. and we do not enter into this here.
A more direct application of the formula in Eq. (9) can be made to discuss the atmospheric neutrino oscillations, which is a vacuum oscillation. However, to get the equivalent of large mixing angle we need to adjust $`\xi 1`$ and $`1/R^2\mathrm{\Delta }m_{atm}^2`$, which means that Eq. (9) is not applicable and we must therefore use Eq. (8) and study its large $`\xi `$ limit.
To discuss this, we start with the limit when $`\xi 1`$. First thing to note is that the pattern of eigenvalues is very different in this case from the small $`\xi `$ case. For small values of $`n`$ the eigenvalues are well approximated by $`\lambda _n=\frac{2n+1}{2}`$, while $`N_n`$ becomes $`n`$ independent until certain cut off value that roughly speaking is given by $`n_\mathrm{\Lambda }=\pi ^2\xi ^2/4`$. Beyond that point $`\lambda _nn`$ and we may no longer neglect the contribution of $`\lambda _n`$ to (6). This results in a suppression of $`N_n`$ that goes like $`1/n`$. It should therefore be reasonable to consider as a first approximation that the expansion (7) is cut off at $`n_\mathrm{\Lambda }`$, and that all the mass eigenstates contributing to $`\nu _L`$ are equally suppressed by $`1/\sqrt{n_\mathrm{\Lambda }}`$. In this approximation, we obtain the survival probability to be
$$P_{surv}(L)=1\frac{2}{n_{\mathrm{\Lambda }}^{}{}_{}{}^{2}}\underset{k,n=0}{\overset{n_\mathrm{\Lambda }}{}}\mathrm{sin}^2\left[\frac{L}{4ER^2}(\lambda _n^2\lambda _k^2)\right].$$
(10)
Certainly, this relation presents a oscillatory profile, however, the superposition of the equally suppressed oscillations will result most of the time on a destructive interference, with the exception of the very sharp resonances that appear each time $`L`$ reaches a multiple of the oscillation length. In other words, $`\nu _L`$ is a superposition of a large number of mass eigenstates, with masses covering a large range, from $`1/2`$ to $`n_\mathrm{\Lambda }`$ in units of $`1/R`$, all of them contributing by the same amount. As a result, once $`\nu _L`$ is released, one may surmise that the time evolution of the different components will most likely wash out the original coherent superposition and the initial $`\nu _L`$ will almost disappear. As fig. 3 shows, this conclusion is borne out by the the numerical analysis. To get an analytical result that also supports this conclusion, note that in the sum in Eq. (10), the dominant contributions come when $`nk`$ in which case each term in the sum averages to $`1/2`$ and on performing the double sum it is easy to see that one arrives at $`\overline{P}_{surv}1/n_\mathrm{\Lambda }`$.
If $`n_\mathrm{\Lambda }`$ is very large, it suppresses the survival probability too much and cannot help in the understanding of the atmospheric data. So, clearly, if we wanted an understanding of the atmospheric data, we must assume smaller $`\xi `$, perhaps values closer to one. In this case, truncating the sum in the expression for the survival probability in Eq. (9), cannot be justified and we must seek an alternative way to deal with Eq. (8). An approach suggested in is to use a continuous approximation to the sum in Eq. (8), which leads to
$$P_{surv}(z)=\left|_0^{\mathrm{}}๐n\frac{e^{izn^2}}{n^2+f(\xi )}\right|^2\xi ^4=\left(\frac{\pi ^2\xi ^4}{4f(\xi )}\right)\left|1\mathrm{erf}\left(\sqrt{izf(\xi )}\right)\right|^2,$$
(11)
where $`z=L/2ER^2`$. Clearly, for large $`\xi `$; $`f(\xi )\pi ^2\xi ^4/4`$, and last expression in Eq. (11) simplifies. In order to study the dependence of $`P_{surv}(z)`$ on $`\xi `$, we plot $`P_{surv}`$ in Fig. 3 in the continuous and discrete approximations. One thing that emerges is the non-oscillatory nature of the function as we exceed $`\xi =1`$. We also see that the continuous limit gives a good approximation for the slope (see figure 3), even for cases where $`\xi 1`$ . This expression, however, does not represent the survival probability in the limit $`\xi 1`$, since in this case (i.e. $`\xi 1`$), the $`P_{surv}`$ has an oscillatory behaviour unlike the last term in Eq. (11) (see figure 4). In the overlap region close to $`\xi 1`$, one may evaluate the second term in Eq. (11) in a slightly different way which leads to an expression for the survival probability as
$$P_{surv}(\zeta )=\rho ^2(\zeta )+\vartheta ^2(\zeta ),$$
(12)
where we have introduced the new variable $`\zeta =\sqrt{zf(\xi )}=(\pi \xi ^2/2R)\sqrt{L/2E}`$ and the functions $`\rho (\zeta )=1C(\zeta )S(\zeta )`$; and $`\vartheta (\zeta )=C(\zeta )S(\zeta )`$ with $`S`$ and $`C`$ the sine and cosine Fresnel integrals:
$`C(\zeta )=\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle _0^\zeta }๐t\mathrm{cos}(t^2)\text{ and }S(\zeta )=\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle _0^\zeta }๐t\mathrm{sin}(t^2).`$We draw attention to the fact that $`\zeta `$ involves not only the model parameters $`\xi `$ and $`R`$ but also the experimental variable $`z`$. Thus for a given model, the variable $`\zeta `$ has different values for different oscillation experiments; for instance, for the atmospheric neutrino case, a typical value for $`\zeta \frac{\pi \xi ^2}{2R}1.2\times 10^2`$ eV<sup>-1</sup>.
The expression (12) is valid just before $`P_{surv}(\zeta )`$ reaches the average, $`\overline{P}_{surv}=4/\pi ^2\xi ^2`$. In figure 5 we present the behaviour of $`\rho `$, $`\vartheta `$ and $`P_{surv}`$. Note the steep fall off of $`P_{surv}`$ as $`\zeta `$ increases. For $`\zeta =1`$, $`P_{surv}`$ has already dropped under 20%, which is smaller than the observed deficit in solar and atmospheric data. Therefore, if we want to fit the overall suppression of the atmospheric neutrinos, we must remain in a very narrow range of parameters. A rough estimate of these parameters may be obtained by expanding (12) to first order on $`\zeta `$, which yields
$$P_{surv}(\zeta )12\sqrt{\frac{2}{\pi }}\zeta =12\sqrt{\pi }\left(\frac{\xi ^2}{R}\right)\left(\frac{L}{4E}\right)^{\frac{1}{2}}.$$
(13)
Now, we may invert the last equation to get
$$\frac{\xi ^2}{R}\frac{(1\overline{P}_{exp})}{\sqrt{4\pi }}\left(\frac{L}{4E}\right)_{exp}^{\frac{1}{2}}.$$
(14)
Using average values for the $`L,E`$ and $`P_{exp}`$ for atmospheric neutrinos, we find $`\xi ^2/R10^2eV`$. Note that this naive approximation gives just about the value obtained by the numerical fitting of the data in . For this value, as $`\xi >1`$, we get $`R`$ in the range $`10^3eV<1/R<10^2eV`$ implying $`1<\xi ^2<10`$. Thus a solution to the atmospheric neutrino puzzle using KK modes requires that $`R`$ be in the sub-millimeter range. This in turn determines how large $`M^{}`$ should be to avoid any unwanted large fine tuning. We see that for $`m10^3eV`$ and Yukawa coupling $`h`$ of order one, we need $`M^{}100TeV`$.
We also point out that in the case of $`\xi 1`$, extra dimensions must be much larger than a millimeter if we want to fit the solar neutrino data either via MSW or via vacuum oscillation. To see this note that if we take $`\mathrm{\Delta }m_{sol}^210^6`$ eV<sup>2</sup>, this would imply $`L/4E1/\mathrm{\Delta }m_{sol}^210^6`$ eV<sup>-2</sup>. Putting this in Eq. (13), we get $`\xi ^2/R10^4`$ eV. For $`\xi 1`$, this implies that $`R0.2`$ mm. Similar estimate for the case of vacuum oscillation yields $`R2`$ cm. Therefore, in all our discussion of solar neutrino oscillations involving the bulk neutrinos, we will work in the approximation $`\xi 1`$.
Let us now summarize our findings for a single neutrino case: for $`\xi 1`$, $`P_{surv}`$ has an oscillatory behaviour as given by (9). As $`\xi `$ approaches 1, $`\lambda _n`$ starts to deviate from the integer value $`n`$, which in turn disturbs the periodic nature of $`P_{surv}`$ and the maximum and minimum values of probability are not reached away from the source . This picture gets worse as one approaches $`\xi 1`$ when a very sharp slope drives the probability near zero as we move away from the source of the neutrinos and it remains around its average, $`4/\pi ^2\xi ^2`$, most of the time. For this reason, whenever we try to fit the atmospheric neutrino data with bulk neutrinos, the value of $`\xi `$ must be tuned to a very narrow range.
Let us apply the discussions of this section to study how the oscillation of known neutrinos to bulk ones would effect the current experiments such as CHOOZ-PALOVERDE, LSND and the atmospheric neutrinos. For this purpose, we first note that in contrast with the usual two neutrino oscillation case, where the pattern is determined by three parameters, the mixing angle $`\theta `$, the $`\mathrm{\Delta }m^2`$ and $`L/4E`$ for the experiment, in the case of bulk neutrino oscillation, we have $`L/4E`$ characterizing an experiment like before, bulk radius $`R`$, which replaces $`\mathrm{\Delta }m^2`$ (which we will assume to be in the milli-meter range) and model parameter $`\xi `$ (which is the analog of the mixing angle). For a given $`R`$, $`\xi `$ is the only parameter characterizing the oscillation pattern. In Fig. 6, we plot the variation of the survival probability $`P_{surv}`$ against $`\xi `$ for the various cases mentioned above for a typical characteristic value of $`L/4E`$. We see from this figure that to explain the observed overall deficit of atmospheric neutrinos by nearly 50%, one needs to go to $`\xi 1.5`$ or so. This conclusion is in acoord with our conclusion based on Eq. (14) above.
## III Three flavour oscillations
Let us now apply the discussion of the previous section to the case of three standard model generations in the brane so that we have three brane neutrinos $`\nu _{e,\mu ,\tau }`$. To give masses to all of them in a minimal scenario, we will use three bulk neutrinos and allow arbitrary Yukawa couplings between the bulk and the brane neutrinos. This leads to an arbitrary Dirac mass matrix that involves the familiar left-handed neutrinos and the right handed Kaluza-Klein modes of the bulk neutrinos. As already discussed in , the most general Dirac mass terms with three flavours may be written, after a rotation of the bulk fields, as
$$=\overline{\nu }_LUM_D\nu _{BR}(y=0)+๐y\overline{\nu }_{BL}_5\nu _{BR}+h.c.,$$
(15)
where $`U`$ is a unitary matrix and $`M_D=Diag(m_1,m_2,m_3)`$, in the basis where $`\nu _L=(\nu _e,\nu _\mu ,\nu _\tau )_L`$; and $`\nu _B=(\nu _B^1,\nu _B^2,\nu _B^3)`$. The mass parameters $`m_\alpha `$ are just the eigenvalues of the Yukawa coupling matrix multiplied by $`v`$, the vacuum expectation value of the standard model doublet field. They are of the order of eV or less since the couplings of the bulk modes to the brane fields are naturally suppressed, as already stated on the previous section.
Now, to simplify the discussion, we rotate the weak eigenstate neutrinos $`\nu _a`$ to the ones related to the weak eigenstates by the rotation $`U`$ i.e. $`\nu _a=U_{a\alpha }\nu _\alpha `$, where $`a=e,\mu ,\tau `$ and $`\alpha =1,2,3`$. We then have
$$=\underset{\alpha =1}{\overset{3}{}}[m_\alpha \overline{\nu }_{\alpha L}\nu _{BR}^\alpha (y=0)+dy\overline{\nu }_{BL}^\alpha _5\nu _{BR}^\alpha +h.c.].$$
(16)
This reduces the problem to a consideration of three KK towers mixing with three neutrinos which are related to the weak interaction eigenstates by the unitary matrix $`U`$ defined above. Each tower is characterized by its $`\xi _\alpha `$ parameter defined in analogy with section II as $`\xi _\alpha \sqrt{2}m_\alpha R`$. After diagonalization of the mass matrices in each tower, each standard neutrino can be written as a coherent superposition of the three different towers of mass eigenstates:
$$\nu _a=\underset{\alpha =1}{\overset{3}{}}U_{a\alpha }\nu _\alpha =\underset{\alpha =1}{\overset{3}{}}\underset{k=0}{\overset{\mathrm{}}{}}U_{a\alpha }\frac{1}{N_{\alpha k}}\stackrel{~}{\nu }_{\alpha k}.$$
(17)
This expression generalizes Eq. (7). It is now clear that the three flavour oscillations will correspond to the oscillations among the three towers. In this regards, the explanation to neutrino puzzles is not any more described in terms of three single neutrinos, as in the usual case; instead all the KK modes can contribute (unless all the KK excitations decouple from the spectrum).
To proceed further, let us define the partial transition probabilities
$$p_{\alpha \beta }(L)\overline{\nu _\alpha (L)|\nu _\alpha (0)}\nu _\beta (L)|\nu _\beta (0).$$
(18)
Notice that the diagonal component $`p_{\alpha \alpha }`$ may be interpreted as the survival probability of $`\nu _\alpha `$, and it takes the form of Eqs. (9) and (12) in that case. It of course involves the $`\nu _\alpha `$โs and not the flavor eigenstates as is obvious. The transition probability among standard flavours can be written in terms of the $`p_{\alpha \beta }`$โs as
$$P_{ab}=\underset{\alpha \beta }{}U_{a\alpha }^{}U_{b\alpha }U_{b\beta }^{}U_{a\beta }p_{\alpha \beta }.$$
(19)
Neglecting all CP phases we may expand Eq. (18) to the form
$$p_{\alpha \beta }=12\underset{k,n=0}{\overset{\mathrm{}}{}}\frac{\mathrm{sin}^2\left(\frac{z}{2}(\lambda _{\alpha n}^2\lambda _{\beta k}^2)\right)}{N_{\alpha n}^2(\xi _\alpha )N_{\beta k}^2(\xi _\beta )}.$$
(20)
Now, we are ready to address the oscillation problem. Our approach will be as follows: we will first select the parameter range that provides overall reduction required to explain the solar and atmospheric data, and then ask whether for the same range of parameters we can explain the observed oscillation between $`\overline{\nu }_\mu `$ to $`\overline{\nu }_e`$ reported by LSND. Without loss of generality, we can assume the hierarchy $`\xi _1<\xi _2<\xi _3`$. We consider three possible scenarios:
* $`\xi _{1,2,3}1`$;
* $`\xi _{1,2}1\xi _3`$ and
* $`\xi _11\xi _{2,3}`$.
The case where all $`\xi _a>1`$ is already ruled out since, as we discussed earlier, it cannot explain the solar neutrino data without implying that the extra dimensions be too large. We therefore do not discuss this case. In the discussion of our results, the following expressions for the partial transition probabilities will be very useful: If $`\xi _{\alpha ,\beta }1`$, then
$$p_{\alpha \beta }=\frac{1}{\eta _\alpha ^2\eta _\beta ^2}\left[\mathrm{cos}\left(\frac{L}{2E}\mathrm{\Delta }m_{\alpha \beta }^2\right)+\left(\xi _\alpha ^2+\xi _\beta ^2\right)c(z)+\xi _\alpha ^2\xi _\beta ^2\left(c^2(z)+s^2(z)\right)\right],$$
(21)
with $`\mathrm{\Delta }m_{\alpha \beta }^2=m_\alpha ^2m_\beta ^2`$; $`\eta _\alpha ^2=1+\pi ^2\xi _\alpha ^2/6`$ and $`z=L/2ER^2`$ as before, and where we have introduced the functions
$$\left(\begin{array}{c}c(z)\\ s(z)\end{array}\right)=\underset{n=1}{\overset{\mathrm{}}{}}\frac{1}{n^2}\left(\begin{array}{c}\mathrm{cos}(zn^2)\\ \mathrm{sin}(zn^2)\end{array}\right).$$
(22)
It is simple to check that the diagonal term of Eq. (21) reduces to Eq. (9). For $`\xi _\alpha 1<\xi _\beta `$ we get
$$p_{\alpha \beta }=\frac{1}{\eta _\alpha ^2}c_\beta (z)+\frac{\xi _\alpha ^2}{\eta _\alpha ^2}\left(c(z)c_\beta (z)+s(z)s_\beta (z)\right);$$
(23)
where we now used $`c_\beta (z)=\mathrm{cos}(\zeta _\beta ^2)\rho (\zeta _\beta )\mathrm{sin}(\zeta _\beta ^2)\vartheta (\zeta _\beta )`$ and $`s_\beta (z)=\mathrm{sin}(\zeta _\beta ^2)\rho (\zeta _\beta )+\mathrm{cos}(\zeta _\beta ^2)\vartheta (\zeta _\beta )`$ with $`\zeta _\beta =\sqrt{z}\pi \xi _\beta ^2/2`$, and $`\rho `$ and $`\vartheta `$ as defined in the previous section, we should stress that all those functions have a similar deep behaviour. And, finally, for $`\xi _{\alpha ,\beta }>1`$ we have
$$p_{\alpha \beta }=c_\alpha (z)c_\beta (z)+s_\alpha (z)s_\beta (z).$$
(24)
Again, it is straightforward to check that the diagonal component of this equation reduces to Eq.(12).
### A Case I: $`\xi _\alpha 1`$
Substituting Eq. (21) into (19), we get, to leading order in $`\xi _\alpha `$
$$P_{ab}=\delta _{ab}2\underset{\alpha ,\beta }{}U_{a\alpha }U_{b\alpha }U_{b\beta }U_{a\beta }\mathrm{sin}^2\left(\frac{L}{4E}\mathrm{\Delta }m_{\alpha \beta }^2\right)+O(\xi _\alpha ^2).$$
(25)
Therefore, as expected, to this order, we obtain the standard expression for the transition probability well known for the three neutrino case. It is clear that in this scenario, solar and atmospheric data can be explained as in the usual three flavour neutrino models by adjusting the spacing of the different $`\xi _\alpha `$โs. For the solar neutrino puzzle, one may either use MSW or VO solution depending on how much fine tuning one is willing to tolerate.
For intermediate values where $`L/4E1/\mathrm{\Delta }m_{\alpha \beta }^2`$, the leading corrections in $`\xi `$ become important. Expanding $`P_{ab}`$ up to order $`\xi ^4`$, by introducing Eq. (21) and neglecting the standard oscillatory term we found
$`P_{ab}`$ $``$ $`\delta _{ab}\left[12{\displaystyle \underset{\alpha }{}}|U_{a\alpha }|^2{\displaystyle \frac{\xi _\alpha ^2}{\eta _\alpha ^2}}\left({\displaystyle \frac{\pi ^2}{6}}c(z)\right)\right]+\left({\displaystyle \underset{\alpha }{}}U_{a\alpha }U_{b\alpha }\xi _\alpha ^2\right)^2I(z);`$ (26)
where we have denoted
$$I(z)=\left[\frac{\pi ^4}{36}+c^2(z)+s^2(z)\frac{\pi ^2}{3}c(z)\right],$$
(27)
Notice that the first term between parenthesis on Eq. (26) contains the lower order correction to the standard survival probability in Eq. (25), while the second term, of oder $`\xi _\alpha ^4`$ will be only relevant for flavour transitions. Taking the average on the last equations, and using that $`\overline{c}=0`$ and $`\overline{c^2+s^2}=\pi ^4/90`$, we get
$$\overline{P}_{ab}\delta _{ab}\left(1\frac{\pi ^2}{3}\underset{\alpha }{}|U_{a\alpha }|^2\frac{\xi _\alpha ^2}{\eta _\alpha ^2}\right)+\frac{7}{180}\pi ^4\left(\underset{\alpha }{}U_{a\alpha }U_{b\alpha }\xi _\alpha ^2\right)^2.$$
(28)
This last expression generalizes that presented in Ref. , where only the contribution of $`\xi _3`$ was assumed.
It is clear that understanding LSND results in this case would require that $`\nu _\mu `$ first undergo a transition to the lower KK modes of the bulk neutrinos and then back to the $`\nu _e`$. One might hope that if we adjusted the extra dimension radius to be small enough $`R^10.22`$ eV, then one would get the right mass difference to fit LSND data. The key question then is to see whether the transition rate comes out right. For this purpose, we consider up to the lowest non-vanishing order in $`\xi _\alpha `$ for $`P_{\mu e}`$ given in (26), which after using the unitarity of $`U`$ and the fact that $`\xi _3^2\xi _1^2=2\mathrm{\Delta }m_{atm}^2R^2`$ turns out to have the form
$$P_{\mu e}\mathrm{sin}^22\theta _{\mu e}\left(\mathrm{\Delta }m_{atm}^2R^2\right)^2I(z)$$
(29)
with $`\mathrm{sin}^22\theta _{\mu e}4(U_{\mu 3}U_{e3})^2`$. It is straightforward to check that $`P_{\mu e}(L=0)=0`$ by using the identities $`c(0)=\pi ^2/6`$ and $`s(0)=0`$. From the limits obtained by the CHOOZ and Palo Verde collaboration, we know that $`|U_{e3}|^2<0.03`$ , and assuming $`|U_{\mu 3}|^2=0.5`$ for maximal mixing, the largest optimistic value for the mixing angle we may take is about $`\mathrm{sin}^22\theta _{\mu e}=0.06`$. By fixing $`L/E`$ as for LSND, $`I(z)`$ becomes only a function of $`R`$ (since $`z=L/2ER^2`$). In figure 7 we have plotted $`I(R)`$ versus $`R`$. From this figure we see that for a reasonable large $`R`$ ($`10eV^1`$), the function $`I(R)`$ has very small values already. The combined effect with the factor $`\mathrm{\Delta }m_{atm}^2R^2<1`$ will reduce $`P_{\mu e}`$ even more. Clearly, $`P_{\mu e}`$ will be maximal for the larger possible value of $`R`$. However, the largest allowed radius that permit us to still be confident in our approach is about $`1/R\sqrt{\mathrm{\Delta }m_{atm}^2}0.06eV`$. A numerical calculation with these inputs gives for LSND $`P_{\mu e}=3\times 10^4`$, which is one order of magnitude smaller than the observed anomaly in the $`\overline{\nu }`$ beam of LSND. Since higher order corrections in $`\xi _\alpha `$ to the probability are unlikely to introduce enough enhancement (a factor of 10 is needed), we conclude that this scenario yields a too small probability for $`\nu _\mu \nu _e`$ transition to explain the LSND observations.
### B Case II: $`\xi _{1,2}1\xi _3`$
To proceed with this case, it is convenient to write the survival probability $`P_{aa}`$ in the following form:
$$P_{aa}=\underset{\alpha ,\beta =1,2}{}\left(U_{a\alpha }U_{a\beta }\right)^2p_{\alpha \beta }+2U_{a3}^2\underset{\alpha =1,2}{}U_{a\alpha }^2p_{\alpha 3}+U_{a3}^4p_{33}.$$
(30)
The first term in the above equation can be written to leading order in the small $`\xi `$โs as follows:
$$\underset{\alpha ,\beta =1,2}{}\left(U_{a\alpha }U_{a\beta }\right)^2p_{\alpha \beta }\left(1U_{a3}^2\right)^2\mathrm{sin}^22\theta _{aa}\mathrm{sin}^2\left(\frac{L}{4E}\mathrm{\Delta }m_{21}^2\right)+O(\xi ^2)$$
(31)
where $`\mathrm{sin}^22\theta _{aa}=4(U_{a1}U_{a2})^2`$. This term by itself can explain the solar neutrino deficit. Clearly, the standard two neutrino oscillation expression is recovered to this order if we set $`U_{e3}0`$ to satisfy the bounds imposed by the reactor data. This will make the contributions of the last two terms in the survival probability (30) negligible and the first oscillatory term can then be used to solve the solar neutrino puzzle. One can of course use the MSW mechanism to solve the solar neutrino problem or the vacuum oscillation. The constraint on the parameter space is that the $`\mathrm{\Delta }m_{12}^2`$ be appropriately adjusted. This does not impose any condition on the bulk radius and can be satisfied by the initial choice of parameters (the Yukawa couplings) in the theory. For instance, $`M^{}10`$ TeV can lead to the MSW-type mass differences. Let us note that, if $`\mathrm{\Delta }m_{12}^210^5`$ eV<sup>2</sup>, then, we will have $`|\xi _1^2\xi _2^2|10^5R^2`$ (eV cm)<sup>2</sup>.
Now let us consider $`P_{\mu \mu }`$. From Eq (30), we see that there are several contributions to the atmospheric neutrino deficit. First, there is the contribution of the towers labeled by $`\xi _{1,2}`$, which is oscillatory, although it can not be identified as in the usual $`\nu _\mu \nu _\tau `$ oscillations. Then, there is also the contribution induced by the term $`p_{33}`$ which is of the form of (12). Finally, there is also a mixed term, $`p_{\alpha 3}`$.
To proceed with the full discussion, let us consider two cases:
Case (i): $`U_{\mu 3}=0`$
If $`U_{\mu 3}=0`$, the last two contributions are removed and we get the survival probability
$$P_{\mu \mu }14\left(2U_{\mu 2}^2\mathrm{\Delta }m_{21}^2R^2+\xi _1^2\right)\left(\frac{\pi ^2}{6}c(z)\right)$$
(32)
One might then hope that the oscillations into the lower KK modes with mass differences of about $`1/R`$ will do the job provided we choose $`1/R\sqrt{\mathrm{\Delta }m_{atm}^2}`$ to match the data. In this case, the atmospheric muon neutrinos oscillate into the sterile neutrinos, a possibility which has its characteristic tests<sup>ยง</sup><sup>ยง</sup>ยงWe realize that from an experimental point of view, this looks less likely to be realized in nature; we take a somewhat liberal view of the situation and still contemplate this as a viable possibility.. For this solution to work, one needs to assume $`\xi _11`$ so that one gets maximal mixing. This in turn means that, to explain both solar and atmospheric neutrino data, an almost degeneracy $`\xi _1\xi _2`$ must be maintained. This alters our explanation of the solar neutrino deficit since, now, these new contribution to it (see the $`O(\xi ^2)`$ terms in Eq. (28)), become more and more important and in fact of order one, making it hard to understand the solar neutrino deficit, since $`c(z)1`$. We will, therefore, consider this case as an unfavorable one for understanding the neutrino puzzles.
Case (ii): $`U_{\mu 3}0`$
Turning to the case where $`U_{\mu 3}0`$, if we keep $`\xi _{1,2}1`$ (to maintain our understanding of the solar neutrino data) then, the corrections of order $`\xi _{1,2}^2`$ to the transition probability become almost negligible and the dominant contribution to $`P_{ab}`$ then must come from $`\xi _3`$ corrections. This yields
$`P_{ab}`$ $``$ $`\left(\delta _{ab}U_{a3}U_{b3}\right)^24(U_{a1}U_{b2})^2\mathrm{sin}^2\left({\displaystyle \frac{L}{4E}}\mathrm{\Delta }m_{21}^2\right)`$ (34)
$`+2U_{a3}U_{b3}\left(\delta _{ab}U_{a3}U_{b3}\right)c_3(z)+(U_{a3}U_{b3})^2p_{33}.`$
Notice that $`p_{33}`$ in the last equation has the same form as Eq. (12), and the function $`c_3`$ is the one defined above with the same sharp behaviour as $`p_{33}`$. Then specializing Eq. (34) to the $`\nu _\mu `$ in the atmospheric case, we get
$$P_{\mu \mu }\left(1U_{\mu 3}^2\right)^2+2U_{\mu 3}^2(1U_{\mu 3}^2)c_3(z)+U_{\mu 3}^4p_{33}.$$
(35)
From our naive analysis in the previous section we may expect that this equation can account for the atmospheric data without too much trouble as long as $`\xi _3^2/R10^2`$ eV or so to make the width of the slope larger than the experimental parameters and to avoid the over washing of the $`\nu _\mu `$ flux. Indeed, it has been checked numerically in reference that the atmospheric neutrino data can be fitted in this case if $`U_{\mu 3}^20.4`$ and $`\xi _3^2/R0.02eV`$. It is worth mentioning that a mixed explanation could be also possible, where the three towers contribute equally to provide atmospheric oscillations, but we will not discuss this case here.
An important point to note however is that due to the features of the function $`c_3(z)`$, the atmospheric neutrino data will not exhibit the oscillatory behaviour that one would expect in the conventional two neutrino models.
Lets turn now to LSND results. By taking that $`U_{e3}0`$ as suggested by solar neutrino and reactor data, the transition probability reduces to $`P_{\mu e}=_{\alpha ,\beta =1,2}U_{\mu \alpha }U_{e\alpha }U_{\mu \beta }U_{e\beta }p_{\alpha \beta }`$. This reflects the fact that the same argument that suppresses the contribution of the third KK tower to $`P_{ee}`$ also does the same for $`P_{\mu e}`$. As $`U_{e3}=0`$ remove the contributions of the third tower, we may expand $`P_{\mu e}`$ to the lowest order by the same expression (26) used in the previous case, which is now given as
$$P_{\mu e}\mathrm{sin}^22\theta _{\mu e}\left(\mathrm{\Delta }m_{sol}^2R^2\right)^2I(z)$$
(36)
where now $`\mathrm{sin}^22\theta _{\mu e}=4(U_{\mu 2}U_{e2})^2`$, and $`I(z)`$ as given in Eq. (27). This resembles our former expression in (29). In order to estimate the magnitude of this contribution, note that the atmospheric neutrino fitting requires $`\xi _3^2/R10^2`$ eV. For $`\xi _3^2110`$, this implies that $`R^210^410^6`$ eV<sup>-2</sup>. Thus if $`\mathrm{\Delta }m_{sol}^2`$ corresponds to the MSW solution (large or small angle), then $`\mathrm{\Delta }m_{sol}^2R^21`$ and one obtains $`P_{\mu e}=\mathrm{sin}^22\theta _{\mu e}I(z)`$. From Fig. 7, we see again that for relevant values of $`R`$ and $`L/E`$, the function $`I(z)`$ takes very small values ($`10^4`$), making the $`P_{e\mu }`$ very small.
Notice that this argument is independent of the way we get the atmospheric deficit. We could also imagine, for instance, that a small value of $`U_{e3}`$ is allowed, and then calculate the leading correction to the above expression. However, it turns out to be of the form $`(U_{\mu 3}U_{e3})_{\alpha =1,2}U_{\mu \alpha }U_{e\alpha }\xi _\alpha ^2(c(z)\frac{\pi ^2}{6})`$, where last term between parenthesis is already smaller than $`10^4`$ by itself.
### C Case III: $`\xi _11\xi _{2,3}`$
In this case, the transition probability can be written as
$$P_{ab}=(U_{a1}U_{b1})^2p_{11}+2U_{a1}U_{b1}\underset{\alpha =2,3}{}U_{a\alpha }U_{b\alpha }p_{\alpha 1}+\underset{\alpha ,\beta =2,3}{}U_{a\alpha }U_{b\alpha }U_{b\beta }U_{a\beta }p_{\alpha \beta }.$$
(37)
The simplest possibility is to let the first term in above equation is be responsible for solar neutrino oscillations into bulk neutrinos as explained by Eq. (9. This requires that the radius be fixed to be about $`1/R10^3eV`$. In this case, to keep the $`\nu _e`$ from mixing too much with the other neutrinos and generate further reduction of the survival probability for the solar neutrino, we choose $`U_{e1}1`$. This essentially decouples the first tower from the others. As a simple approximation, if we assume that $`U_{e,2,3}=0`$, then clearly this suppresses the oscillations from $`\nu _\mu `$ into $`\nu _e`$, making it difficult to understand the LSND results.
Moreover, in this scenario, even the explanation for atmospheric data seems to run into some trouble. As $`U_{e1}1`$, it is not unreasonable to conclude based on orthogonality that $`U_{\mu 1}0`$. This implies that
$$P_{\mu \mu }=\underset{\alpha ,\beta =2,3}{}(U_{\mu \alpha }U_{\mu \beta })^2p_{\alpha \beta },$$
(38)
where all $`p_{\alpha \beta }`$ are given as in Eq. (24). As a rough approximation, if we assume that the partial transition probabilities are all almost of the same order (as they seem to be numerically), say $`p_{33}`$, and use orthogonality once more to get $`P_{\mu \mu }p_{33}`$. Therefore, we get maximal contribution from the large deficit generated by $`p_{33}`$. A hierarchical $`p_{\alpha \beta }`$ will not help with this, since either we get equal mixing among them or one of them has a dominant contribution. In any case, the leading order will combine both things, a large and fast developing slope, and a large mixing angle. Nevertheless, based on the results of , where they found some scenarios where those two ingredients come together (although for small $`\xi `$), it is still hard to rule out the scenario without a careful numerical analysis of the data. In any case, the simplest requirement to get a deficit not too large compared with the experimental data, imposes a tuning of the main parameters $`\xi _{2,3}^2/R10^2`$, which combined with the condition for $`R`$ to understand the solar neutrino data fixes $`\xi _{2,3}^210`$.
In order to understand the LSND results, we must allow for small $`U_{e2}`$ and/or $`U_{\mu 1}`$. Since, in the present scenario, the solar neutrino deficit is being explained purely by oscillation into the bulk, so, we get the following picture: The first tower contributions are too small for LSND, they are even smaller than the size of those considered on the previous cases, while the oscillations involving the other towers are mainly destructive due to large values of $`\xi `$ involved. The conversion will not occur but for a small contribution related with the small part of those towers contained in $`\nu _e`$. However, as one may suspect, the contributions are all proportional to the mixing angles $`(U_{\mu \alpha }U_{e\alpha })^2`$, which can not be larger than $`10^2`$. So, the question is whether the behaviour of $`p_{\alpha \beta }`$ can help. To analyze this point we notice that
$$P_{\mu e}\underset{\alpha ,\beta =2,3}{}U_{\mu \alpha }U_{e\alpha }U_{e\beta }U_{\mu \beta }(1p_{\alpha \beta }).$$
(39)
Note that one can estimate the value of $`(1p_{\alpha \beta })`$ from the the information on atmospheric neutrino deficit as follows. First point is that all $`p_{\alpha \beta }`$โs are of same order and therefore, one can write the $`P_{\mu e}(U_{\mu 1}U_{e1})^2(1P_{surv})`$, where $`P_{surv}(\zeta )`$ is the same function which for $`\zeta =\zeta _{atmos}`$ gives the survival probability of muon neutrinos in the atmospheric data. Thus, $`P_{surv}(\zeta _{atm})0.5`$. The value of $`\zeta `$ corresponding to the LSND case is however much smaller due to smaller oscillation distance; therefore to estimate $`P_{\mu e}`$ we must use the value of $`\zeta `$ which is much smaller and find the corresponding $`P_{surv}`$. A numerical evaluation gives $`(1P_{surv})10^2`$. The next question is how large the mixing parameters are. In order not to make $`P_{ee}`$ in atmospheric neutrino data different from one, $`U_{\mu 2}`$ must be much smaller than 1. On the other hand to fit LSND observations, we need to have $`U_{\mu 1}0.5`$. It therefore appears difficult to accommodate the LSND observations in this case.
## IV Concluding remarks
The analysis of the present paper shows that in minimal models for neutrino masses in theories with large extra dimensions, it is not possible to get a simultaneous explanation of gross overall oscillations needed to understand the solar, atmospheric and the LSND data. Fitting the overall deficit in the atmospheric and solar neutrino data for various ranges of the input parameters, the largest conversion probability, $`P_{\mu e}`$, that we find, is around $`10^4`$. This is too low to explain the LSND observations. One must therefore invoke new physics beyond the standard model in the brane or new physics outside the brane for a simultaneous understanding of all observed neutrino data.
The oscillation pattern is governed by the dimensionless parameters $`\xi _\alpha `$. Three of them arise in the models under consideration and we get the following picture:
* $`\xi _{1,2,3}1`$: in this case, solar and atmospheric neutrino data are understood as in the case of four dimensional models and the hope that the presence of the bulk neutrinos with an appropriate KK excitation scale could provide an understanding of the LSND data is not realized.
* $`\xi _{1,2}1\xi _3`$: solar neutrino data is provided as in the four dimensional models but atmospheric data is explained by $`\nu _\mu `$ to $`\nu _{bulk}`$ oscillation once the appropriate values of the bulk radius is chosen i.e. either $`\xi _1\xi _2`$ and $`1/R\sqrt{\mathrm{\Delta }m_{atm}^2}`$ or $`\xi _3^2/R10^2`$. In this case, the atmospheric neutrino flux does not oscillate as a function of $`L/E`$, a feature distinct from the conventional two neutrino oscillation picture.
* $`\xi _11\xi _{2,3}`$. Both, solar and atmospheric data are explained by $`\nu \nu _{bulk}`$ oscillations. Therefore $`1/R^2\mathrm{\Delta }m_{sol}^210^3eV`$ with matter effects for solar and $`\xi _{2,3}^210`$. Again, the $`L/E`$ behaviour of the atmospheric flux is not oscillatory.
* $`\xi _{1,2,3}1`$. There is no explanation for solar neutrino data in this case. This parameter range is therefore ruled out.
Acknowledgements. The work of RNM is supported by a grant from the National Science Foundation under grant number PHY-9802551. The work of APL is supported in part by CONACyT (Mรฉxico).
|
warning/0006/hep-ph0006198.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The latest LEP bound on the Standard Model like Higgs boson mass $`m_h`$ is in the region of $`m_h>108`$ GeV once the limits from all the experiments are combined . Such a bound on the Higgs mass can be interpreted as the lightest scalar Higgs mass bound of $`m_h>105`$ GeV in the low $`\mathrm{tan}\beta (4)`$ region of the minimal supersymmetric standard model (MSSM). Such a large Higgs mass implies large fine-tuning in the low $`\mathrm{tan}\beta `$ region , and leads to the smallest values of $`\mathrm{tan}\beta `$ being excluded. This is disturbing since electroweak baryogenesis in the MSSM relies on low $`\mathrm{tan}\beta `$ and light Higgs and stop masses <sup>1</sup><sup>1</sup>1When SUSY phases are included the parameter space for electroweak baryogenesis is somewhat increased, however ..
It is well known that in the next-to-minimal supersymmetric standard model (NMSSM) the lightest Higgs boson can be heavier than in the MSSM . It is also well known that the parameter space for electroweak baryogenesis in the NMSSM is much larger than in the MSSM, due to trilinear contributions to the scalar potential at tree-level. What has not been realised so clearly is that these two features go together with low $`\mathrm{tan}\beta `$. This is a timely observation, given the constraints that LEP places on the MSSM in the low $`\mathrm{tan}\beta `$ region. Moreover, although fine-tuning has been well studied in the MSSM it has not been systematically studied in the NMSSM. In this letter, then, we compare the fine-tuning in the NMSSM and MSSM in the low $`\mathrm{tan}\beta `$ region and show that the NMSSM is much preferred. We then study the strength of the electroweak phase transition (EWPT) in the NMSSM, and show that, unlike in the MSSM, a sufficiently strong first order phase transition persists over much of the low $`\mathrm{tan}\beta `$ region.
Although the NMSSM provides a testable solution to the $`\mu `$ problem, by replacing the superpotential term $`\mu H_1H_2`$ by $`\lambda NH_1H_2`$ (where $`H_i`$ are Higgs doublets and $`N`$ is a Higgs singlet) it is often criticised for leading to a conflict between cosmological domain walls and stability due to supergravity tadpoles. It has recently been pointed out, however, that the NMSSM remains a natural solution to the $`\mu `$ problem since both the stablility and the cosmological domain wall problems may be eliminated by imposing a $`\text{ZZ}_2`$ R-symmetry on the non-renormalisable operators . Thus the NMSSM appears to be well motivated both theoretically and phenomenologically at the present time. Indeed the solution of the $`\mu `$ problem is closely linked to the two phenomenological features of NMSSM noted above in the sense that all the three originate from the same term in the superpotential, as we shall see below.
## 2 Fine-Tuning in the NMSSM
Although fine-tuning is not a well defined concept, the general notion of fine-tuning is unavoidable since it is the existence of fine-tuning problem in the standard model which provides the strongest motivation for low energy supersymmetry, and the widespread belief that superpartners should be found before or at the LHC. If one abandons the notion of fine-tuning then there is no reason to expect superpartners at the LHC. Although a precise measure of absolute fine-tuning is impossible, the idea of relative fine-tuning can be helpful in selecting certain models and regions of parameter space over others. It is useful to compare different models using a common definition of fine-tuning
$$\mathrm{\Delta }_a=abs\left(\frac{a}{M_Z^2}\frac{M_Z^2}{a}\right)$$
(1)
where $`a`$ is an input parameter, and fine-tuning $`\mathrm{\Delta }^{max}`$ is defined to be the maximum of all the $`\mathrm{\Delta }_a`$. It is worth pointing out that at low values of $`\mathrm{tan}\beta `$ fine-tuning is worse in the MSSM for two separate reasons. First, the tree level contribution to the Higgs mass squared upper bound $`M_Z^2\mathrm{cos}^22\beta `$ goes down; so that one must rely more on radiative corrections to meet the LEP bound, which demand a higher value of $`M_3(0)`$ as observed in . Second, the top quark Yukawa coupling is larger for low values of $`\mathrm{tan}\beta `$, so the Higgs mass gets driven more negative, resulting in larger coefficients of the $`M_3^2(0)`$ term in Eq. (1), which again increases fine-tuning. A quantitative study of fine-tuning reveals that it is the experimental limit on the Higgs mass rather than the gluino mass, that provides the most severe fine-tuning for low values of $`\mathrm{tan}\beta `$, as discussed in the second reference in . Moreover the fine-tuning required in the MSSM increases exponentially with the Higgs mass, since the radiative corrections to the Higgs mass increase logarithmically with the stop masses, and the stop masses increase in proportion to $`M_3(0)`$ which controls the fine-tuning .
While the Higgs quartic coupling in the MSSM is related to the gauge couplings, in the NMSSM it has an additional contribution from the Yukawa coupling $`\lambda `$. Consequently the tree level mass limit of the lightest Higgs scalar gets modified to
$$m_h^2M_Z^2\left(\mathrm{cos}^22\beta +\frac{2\lambda ^2}{g^2+g^2}\mathrm{sin}^22\beta \right).$$
(2)
Note that the additional term in Eq. (2) is most effective where its help is needed most, i.e. at low $`\mathrm{tan}\beta `$. Assuming this Yukawa coupling to remain perturbative up to the unification scale of $`10^{16}`$ GeV implies ฮป
<
0.7
<
๐0.7\lambda\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}0.7 at the electroweak scale , i.e. a contribution of $`M_Z^2\mathrm{sin}^22\beta `$ to the tree level mass limit. On the other hand, the upper limit on the mass of the lightest CP even Higgs is not necessarily physically relevant, since its coupling to the $`Z`$ boson can be very small. Actually, this phenomenon can also appear in the MSSM, if $`\mathrm{sin}^2(\beta \alpha )`$ is small. However, the CP odd Higgs boson $`A`$ is then necessarily light ($`m_Am_h<M_Z`$ at tree level), and the process $`ZhA`$ can be used to cover this region of the parameter space in the MSSM. In the NMSSM, a small gauge boson coupling of the lightest CP even Higgs is usually related to a large singlet component, in which case no (strongly coupled) light CP odd Higgs boson is available. Hence, Higgs searches in the NMSSM have possibly to rely on the search for the second lightest Higgs scalar, which can be substantialy heavier than the limit given in Eq. (2) . In the limit that the singlet decouples, and is the heaviest scalar, the inequality in Eq. (2) may be saturated by the lightest Higgs boson. Alternatively if the singlet decouples, but is the lightest scalar, then the inequality in Eq. (2) may be saturated by the second lightest scalar, which is the physical Higgs boson. In fact this latter case approximately applies to the results we present later in the figures. In all cases it is clear that the NMSSM does not require a large radiative correction to satisfy the Higgs mass limit from LEP. It is this modification of the Higgs sector in the NMSSM which is responsible for the opening up the low $`\mathrm{tan}\beta `$ region of parameter space in this model, by removing the connection between the Higgs mass and exponential increases in fine-tuning present in the MSSM.
The NMSSM superpotential is defined as
$$W=\lambda NH_1H_2+\frac{k}{3}N^3+\mathrm{}$$
(3)
where the elipsis stand for quark and lepton Yukawa couplings, and the Higgs potential is
$`V_{NMSSM}`$ $`=`$ $`m_1^2v_1^2+m_2^2v_2^22m_3^2v_1v_2+\lambda ^2v_1^2v_2^2`$ (4)
$`+`$ $`{\displaystyle \frac{1}{8}}(g^2+g^2)(v_1^2v_2^2)^2+x^2(m_N^2+{\displaystyle \frac{2}{3}}kxA_k+k^2x^2)`$
where $`v_i=H_i`$, $`x=N`$ and
$$m_1^2=m_{H_1}^2+\lambda ^2x^2,m_2^2=m_{H_2}^2+\lambda ^2x^2,m_3^2=\lambda x(kx+A_\lambda ).$$
(5)
$`A_\lambda ,A_k`$ are the trilinear soft parameters associated with the $`\lambda ,k`$ terms in the superpotential, which play a prominent role in ensuring a strong 1st order EWPT, as we shall see in the next section. One-loop corrections to the effective potential, $`\mathrm{\Delta }V^{(1)}`$, are taken into account by redefining the scalar masses such that $`m_i^2m_i^2+\mathrm{\Delta }V^{(1)}/v_i^2`$. The NMSSM minimisation conditions are then
$`{\displaystyle \frac{M_Z^2}{2}}`$ $`=`$ $`{\displaystyle \frac{m_1^2m_2^2\mathrm{tan}^2\beta }{\mathrm{tan}^2\beta 1}}`$
$`\mathrm{sin}2\beta `$ $`=`$ $`{\displaystyle \frac{2m_3^2}{m_1^2+m_2^2+\lambda ^2v^2}}`$ (6)
$`0`$ $`=`$ $`\lambda x(2kx+A_\lambda ){\displaystyle \frac{v_1v_2}{x^2}}+kx(2kx+A_k)+(\lambda ^2v^2+m_N^2)`$
The results in Eq. (6) can then be used to derive a master formula for the derivative of the $`Z`$ mass with respect to input parameters, from which one can obtain the sensitivity parameter in Eq. (1)
$`{\displaystyle \frac{M_Z^2}{a}}`$ $`=`$ $`{\displaystyle \frac{2}{\mathrm{tan}^2\beta 1}}\{{\displaystyle \frac{m_1^2}{a}}\mathrm{tan}^2\beta {\displaystyle \frac{m_2^2}{a}}{\displaystyle \frac{\mathrm{tan}\beta }{\mathrm{cos}2\beta }}(1+{\displaystyle \frac{M_Z^2\lambda ^2v^2}{m_1^2+m_2^2+\lambda ^2v^2}})`$ (7)
$`\times `$ $`[2{\displaystyle \frac{m_3^2}{a}}\mathrm{sin}2\beta ({\displaystyle \frac{m_1^2}{a}}+{\displaystyle \frac{m_2^2}{a}}+{\displaystyle \frac{\lambda ^2v^2}{a}})]\}.`$
This result is very similar to the master formula obtained in the MSSM, to which it reduces in the limit $`\lambda 0`$. Of course the partial derivatives on the right-hand side will be quite different since they bring in derivatives of $`\lambda x`$ rather than $`\mu `$, and $`x`$ is a function of all the soft parameters (unlike $`\mu `$ in the MSSM which is independent of the soft parameters.) Thus the implementation of the master formula in the NMSSM is more involved than in the MSSM. Nevertheless, the variation of the VEVs follows from the minimisation conditions of the potential, such that:
$$\frac{^2V_{NMSSM}}{av_i^2}=0,\frac{^2V_{NMSSM}}{ax^2}=0.$$
(8)
The above system of 3 coupled equations can then be solved for $`v_i/a`$, and $`x/a`$. The other partial derivatives we need to evaluate the master formula Eq. (7), such as $`\lambda /a`$, $`m_{H_i}^2/a\mathrm{}`$, are computed numerically when running the renormalisation group equations (RGEs).
The NMSSM minimisation conditions are clearly analagous to those of the MSSM, and they may be satisfied by an analagous proceedure. In both models the soft parameters are chosen at the high energy scale. Once the (low energy) value of $`\mathrm{tan}\beta `$ is selected, the top mass fixes the low energy top Yukawa coupling<sup>2</sup><sup>2</sup>2We include only QCD corrections when converting pole mass to running mass at the $`m_Z`$ scale.. In the NMSSM the additional low energy values of $`\lambda `$ and $`k`$ are selected. Next, all the Yukawas are run up to the high energy scale. Then all the soft masses are run down to low energies and the value of $`x`$ (the analogue of $`\mu `$ in the MSSM) is fixed by the first minimisation condition (up to an ambiguity in the signs), and the value of $`A_\lambda `$ (the analogue of $`B`$ in the MSSM) is fixed by the second minimisation condition. In the NMSSM it then only remains to satisfy the third minimisation condition. In general this requires an iterative process since it relies upon having chosen the correct values of $`A_k`$ and $`m_N`$ at high energies (and both these parameters must be fixed before all the soft masses can be run down). <sup>3</sup><sup>3</sup>3In the limit that we neglect the $`\lambda `$ and $`k`$ contributions to the right-hand sides of RGEs, the RGEs for the soft masses $`m_Q^2,m_U^2,m_{H_1}^2,m_{H_2}^2`$ become identical to those in the MSSM, and that for $`A_\lambda `$ becomes identical to that for the soft parameter $`B`$ in the MSSM. Furthermore $`m_N`$, $`A_k`$ and $`k`$ do not run in this limit. In this limit the first two minimisation conditions are identical to those of the MSSM and the third can trivially be satisfied by using it to fix $`m_N`$ at the end. Although we never make this approximation, it serves to emphasise that main differences between the NMSSM and the MSSM arise from the low energy Higgs potential and scalar mass squared matrix not from RG running. The structure of the Higgs potential being more complicated in the NMSSM than in the MSSM, the minimisation conditions (6) do not guarantee that we sit at a local minimum rather than at a local maximum, in which case the lightest squared mass eigenvalue is negative. The acceptable regions of the low energy parameter space where the physical scalar squared masses are positive were mapped out by Elliott et al in , and may be straightforwardly be achieved by adjusting the high energy soft Higgs masses $`m_{H_1}(0)`$ and $`m_{H_2}(0)`$. We therefore take a common value $`m_0`$ at $`M_{GUT}`$ for the squark and slepton masses, but allow the input Higgs masses to be different. Also, a positive squared Higgs mass spectrum is favoured when $`A_k(M_Z)`$ is small, so we have adjusted the initial value of $`A_k(0)`$ such that $`A_k(M_Z)=0`$.
In figs. 1-3 we plot the maximum sensitivity parameter $`\mathrm{\Delta }^{max}`$ as a function of the lightest physical Higgs mass for both the NMSSM and MSSM, for the values of $`\mathrm{tan}\beta =2,\mathrm{\hspace{0.17em}3},\mathrm{\hspace{0.17em}5}`$. In these plots we have taken $`m_0=100`$ GeV, $`A_t(0)=0`$ GeV, and assumed high energy first and second gaugino massses given by $`M_2(0)=M_1(0)=500`$ GeV. In our scans over parameter space we have restricted ourselves to $`100\mathrm{GeV}<M_3(0)<600\mathrm{GeV}`$, $`0<m_{H_1}(0)<1\mathrm{TeV}`$, and $`\mu <0`$ ($`\lambda x<0`$). In the MSSM, for each pair of points ($`M_3(0),m_{H_1}(0)`$) the upper limit on $`m_{H_2}(0)`$ is set by demanding the lightest chargino to be heavier than 90 GeV. For the NMSSM, we have taken $`\lambda (0)=1`$ and $`k(0)=0.1`$ as a sample point. The values of $`A_\lambda (0)`$ and $`m_N(0)`$ are obtained from the minimisation conditions, and we have restricted the range in $`m_{H_2}(0)`$ to those values which give rise to a physical Higgs spectrum not excluded by LEP. For the three values of $`\mathrm{tan}\beta `$ shown in the plots, the physical Higgs mass plotted is the second lightest Higgs, while the lightest one is a dominantly singlet Higgs with a very weak coupling to the $`Z`$ boson. In fact, a weakly coupled Higgs as light as a few GeV might have escaped detection at LEP<sup>4</sup><sup>4</sup>4An analytical fit to the experimental results giving the constraint between the mass of the scalar Higgs and its coupling to the $`Z`$ boson can be found in .. In both models, MSSM and NMSSM, the increase of the maximum fine-tuning when increasing the physical Higgs mass is mainly due to the increase in $`M_3(0)`$ . Varying $`m_{H_1}(0)`$ and $`m_{H_2}(0)`$ has almost no effect in fine-tuning in the MSSM for values of $`\mathrm{tan}\beta 3`$, and the regions obtained are narrower than in the NMSSM.
It is clear that for all three values of $`\mathrm{tan}\beta `$ the physical Higgs boson mass in the NMSSM may be heavier than in the MSSM, with correspondingly lower fine-tuning. The effect is clearly greater for lower values of $`\mathrm{tan}\beta 3`$, where the experimentally allowed values of the lightest Higgs mass would imply large fine-tuning in the MSSM, whereas in the NMSSM we can have a large enough Higgs mass with a relatively low $`\mathrm{\Delta }^{max}`$. The plots are a striking demonstration that the physical Higgs boson can be heavier and involve less fine-tuning in the NMSSM compared to the MSSM at low values of $`\mathrm{tan}\beta `$.
## 3 Baryogenesis in the NMSSM
Having shown that the low $`\mathrm{tan}\beta `$ region of the NMSSM permits a heavier physical Higgs boson consistent with the LEP limit, without large fine-tuning, we now turn to the question of electroweak baryogenesis in the NMSSM. The anomaly mediated electroweak processes are known to provide an efficient mechanism for baryogenesis in the symmetric phase . In order to prevent the washout of the resulting baryon asymmetry by the back reaction, however, one requires a strongly 1st order EWPT,
$$v_c>T_c,$$
(9)
where $`T_c`$ is the critical temperature and $`v_c`$ the Higgs VEV in the symmetry breaking phase at this temperature. This imposes serious constraints on the underlying model, since it requires a negative cubic term in the generic Higgs potential
$$V=m^2\varphi ^2\mu \varphi ^3+\lambda \varphi ^4,$$
(10)
with
$$v_c=\mu /2\lambda >T_c,$$
(11)
where the critical point is defined as the point of degeneracy between the symmetric and symmetry breaking vacua.
In the SM, thermal loops provide a positive quadratic term as well as a negative cubic term to (10), the latter coming from the $`W`$ and $`Z`$ boson loops. Being a loop effect however the resulting $`\mu `$ is small compared to $`T_c`$, which is typically of the electroweak scale. This implies a stringent limit on $`\lambda `$ and hence on the SM Higgs mass, mh
<
40
<
subscript๐40m_{h}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}40 GeV , which is ruled out by the LEP data. In the MSSM the cubic term gets an additional contribution from stop loops; so that for a light stop $`(m_{\stackrel{~}{t}_R}100\mathrm{GeV})`$ the Higgs mass limit goes up to mh
<
100
<
subscript๐100m_{h}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}100 GeV for $`m_AM_Z`$ . But the twin requirements on $`m_h`$ and $`m_A`$ squeezes the MSSM solution to the low $`\mathrm{tan}\beta `$ region, which is disfavoured by LEP. Indeed the only way of reconciling the LEP limit with the low $`\mathrm{tan}\beta `$ region is by invoking large stop mass and mixing parameters, in which case one can not satisfy (9). Thus the MSSM scenario for a strong EWPT (9) is strongly disfavoured by LEP data, if not ruled out by it altogether.
As already noted in , it is much easier to satisfy the strong EWPT requirement (9) in the NMSSM compared to the MSSM, since the tree level potential (4) itself contains a cubic term $`2\lambda A_\lambda xv_1v_2`$. We closely follow the approach of in using the simplest $`\text{ZZ}_3`$ symmetric form of the NMSSM superpotential (3) and simply retaining the leading term in the expansion of the thermal loop in $`M/T`$, i.e. the mass of the exchanged particle relative to the temperature. Thus
$$V=V_0+V_T,$$
(12)
where $`V_0`$ is the zero-temperature potential represented by (4) along with the radiative correction from top/stop loops, and
$`V_T`$ $`=`$ $`{\displaystyle \frac{T^2}{24}}[\mathrm{Tr}M_S^2+\mathrm{Tr}M_P^2+2m_{H^+}^2+6M_W^2+3M_Z^2`$ (13)
$`+6m_t^2+2m_C^2+m_{N1}^2+m_{N2}^2+m_{N3}^2].`$
The last four terms represent the charged and the neutral Higgsinos. The other superparticle masses are assumed to be larger than $`2T`$ and hence suppressed by the Boltzmaan factor . We shall evaluate the field-dependent masses of (13) in the Landau gauge as in instead of the Unitary gauge of ref. , in view of the well-known ambiguity in calculating finite temperature effects in the latter (the so-called unitary gauge puzzle ). This gives
$`V_T`$ $`=`$ $`{\displaystyle \frac{T^2}{24}}[4m_{H_1}^2+4m_{H_2}^2+2m_N^2+2(3g^2+g^2+3\lambda ^2)(v_1^2+v_2^2)`$ (14)
$`+6h_t^2v_2^2+12(\lambda ^2+k^2)x^2].`$
To get the potential as a function of the Higgs fields $`H_1,H_2,N`$ one has to simply substitute these quantities for their VEVs $`v_1,v_2,x`$ in Eqs. (4) and (14).
For a given $`\mathrm{tan}\beta `$ and stop mass and mixing parameters we first minimize the $`T=0`$ potential (4) along with the radiative correction and impose the experimental constraints on the output Higgs parameters to find an appropriate set of $`\lambda ,k,A_\lambda ,A_k`$ and $`x`$ parameters. We then find the $`T_c`$ as the largest $`T`$ at which the curvature (second derivative) of the potential changes sign along any direction at the origin. It is clear from Eqs. (4), (5), (12) and (14) that
$$T_c^2=\mathrm{Max}[\frac{12m_{H_1}^2}{3g^2+g^2+3\lambda ^2},\frac{12(m_{H_2}^2+\mathrm{\Delta }_{\mathrm{rad}})}{3g^2+g^2+3\lambda ^2+3h_t^2},\frac{2m_N^2}{\lambda ^2+k^2}].$$
(15)
where $`\mathrm{\Delta }_{\mathrm{rad}}=\frac{3h_t^2}{8\pi ^2}m_t^2\mathrm{}n\frac{m_{\stackrel{~}{t}}}{m_t}`$ along with a stop mixing contribution. Although the $`T_c`$ determined from this saddle point of the potential at the origin is not identical to the one determined by the above mentioned degeneracy between the two vacua, they have been shown to agree within 2% . Therefore we use this $`T_c`$ and minimise the potential at this temperature to determine the corresponding VEVs<sup>5</sup><sup>5</sup>5Our minimisation code for the NMSSM potential at finite termperature is based on the corresponding code of Manuel Drees for the zero temperature case.. Since both the doublet Higgs fields contribute to the sphaleron energy the relevant VEV for the requirement (9) is
$$v_c=\sqrt{2}(v_{1c}^2+v_{2c}^2)^{\frac{1}{2}},$$
(16)
where the factor $`\sqrt{2}`$ is due to the normalisation convention, $`M_Z^2=(g^2+g^2)(v_1^2+v_2^2)/2`$ as emphasised in .
The stop mass and mixing parameters are fixed at
$$m_{\stackrel{~}{t}}=500\mathrm{GeV},A_t=1000\mathrm{GeV},$$
(17)
and input values of $`\mathrm{tan}\beta `$ are taken as $`2,3,5`$ and $`10`$. <sup>6</sup><sup>6</sup>6We obtain similar results for $`m_{\stackrel{~}{t}}=250\mathrm{GeV}`$, $`A_t=500\mathrm{GeV}`$. For each $`\mathrm{tan}\beta `$ we run the program at $`7\times 10^5`$ points, corresponding to 7 different starting points in the parameter space. Only about 0.1% of these points give solutions satisfying the experimental constraints along with the requirement that in each case the solution represents the absolute minimum of the potential. This is about an order of magnitude less than the passing rate of ref. , which could be due to our using the most constrained ($`\text{ZZ}_3`$ symmetric) form of the NMSSM as well as the stronger experimental constraints coming from LEP now. About 10% of these solutions satisfy the strong EWPT criterion (9). Thus for each of the 7 starting points in the parameter space we have
<
10
<
absent10\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}10 solutions satisfying all these criteria.
Table 1 shows two representative solutions for each $`\mathrm{tan}\beta `$ along with the corresponding parameter values as well as the output Higgs boson masses, $`v_c`$ and $`T_c`$. One of the solutions corresponds to a relatively low value of the lightest scalar mass, escaping the LEP constraint because of its large ($`98\%`$) singlet content, while the other has a lightest scalar mass above the LEP limit. We see from this table that it is possible to get acceptable Higgs mass spectra as well as satisfy $`v_c>T_c`$ with reasonable values of the parameters, at least for low values of $`\mathrm{tan}\beta `$ (2,3,5). Interestingly one gets acceptable solutions for $`\mathrm{tan}\beta =10`$ as well; but in this case $`k>\lambda `$. The reason is that for large $`\mathrm{tan}\beta `$ the 3rd and 4th terms of the potential (4) become vanishingly small. The latter implies that the $`m_h^2`$ limit of Eq. (2) is dominated by the 1st term, which is however quite large now; while the former implies that the dominant cubic term of the potential is $`\frac{2}{3}kA_kx^3`$. Thus the NMSSM can help to get a strong EWPT in the large $`\mathrm{tan}\beta `$ region as well, where it has little effect on the Higgs mass limit of the MSSM.
It is appropriate to briefly discuss here the role of NMSSM in generating sufficient amount of CP asymmetry, as required for baryogenesis . Of course quantitative investigation of this question depends on the model of electroweak baryogenesis, which is beyond the scope of this work. But it is generally agreed that the size of CP violation in the SM, arising from the complex CKM matrix, is much too small as it is suppressed by the small Yukawa couplings as well as the CKM mixing angles. There are additional sources of CP violation in the MSSM, arising from the phases of $`\mu `$ and the SUSY breaking terms, which can serve this purpose provided the size of the phase angles are $`O(10^1)`$ . On the other hand the experimental constraint from the electric dipole moments of neutron and electron would require these phase angles to be $`O(10^2)`$ unless there is a systematic cancellation between them or one assumes the sfermions of the 1st two generations to have masses $`1`$ TeV . This potential conflict with the electric dipole moment limits is alleviated in the NMSSM, where the required size of phase angles is an order of magnitude smaller than in MSSM . Even more interestingly the NMSSM offers the possibility of generating a spontaneous CP violation in the symmetric phase, which goes down to zero in the symmetry breaking phase . Thus it can effectively contribute to baryogenesis, which takes place in the symmetric phase, while making no contribution to the measured electric dipole moments. It is not possible to generate such a transitional CP violation in the MSSM .
## 4 Conclusion
The current LEP limit on the lightest Higgs boson mass places severe constraints on the MSSM in the low $`\mathrm{tan}\beta `$ region, which was the favoured region for a strong EWPT as required for electroweak baryogenesis. The only way to escape this $`m_h`$ limit is to invoke very large stop mass and mixing parameters, which would imply however large fine-tuning as well as a weak EWPT. Thus one has to sacrifice the naturalness of electroweak symmetry breaking as well as baryogenesis. However both these problems can be solved simultaneously with the so called $`\mu `$ problem of the MSSM by going to the NMSSM. All the relevant terms โ i.e. $`\lambda ^2x^2v_{1(2)}^2`$, $`\lambda ^2v_1^2v_2^2`$ and $`\lambda A_\lambda xv_1v_2`$ โ originate from the superpotential $`\lambda NH_1H_2`$. While the 1st term solves the so-called $`\mu `$ problem, the 2nd provides an additional contribution to the tree-level $`m_h`$ limit and the 3rd one a cubic term in the tree-level potential. Thus the 2nd term alleviates the fine-tuning problem arising from the $`m_h`$ limit, while the 3rd ensures a strong EWPT.
It may be noted here that both the 2nd and the 3rd terms vanish at large $`\mathrm{tan}\beta `$. Thus it does not affect the $`m_h`$ limit of the MSSM at large $`\mathrm{tan}\beta `$, which is any way quite high. However in this case the soft cubic term $`kA_kx^3`$ helps to generate a strong EWPT. Thus the NMSSM helps to give a strong EWPT along with solving the $`\mu `$ problem even in the large $`\mathrm{tan}\beta `$ region.
To summarise, the LEP limit on the Higgs boson mass severely constrains the low $`\mathrm{tan}\beta `$ region of the MSSM, leading to large fine-tuning and problems with electroweak baryogenesis. We have shown that in the low $`\mathrm{tan}\beta `$ region the NMSSM is in much better shape phenomenologically, since the physical Higgs boson masses are larger, the fine-tuning is less, and the electroweak phase transition is more strongly first order in the NMSSM, as compared to the MSSM.
Finally we remark that although we have considered the NMSSM for simplicity, we would expect similar effects in more complicated extensions of the NMSSM, for example those involving an additional anomalous $`U(1)`$ gauge group , or those including Higgs triplets , since both models also involve the $`\lambda NH_1H_2`$ coupling considered here.
Acknowledgements
The authors would like to thank the organisers of WHEPP-6, where this project was started, and B. Ananthanarayan and P.N. Pandita for their contributions during the initial stage of the work. We would also like to thank C.D. Froggatt and M. Pietroni for several helpful communications.
|
warning/0006/astro-ph0006074.html
|
ar5iv
|
text
|
# Dark matter phase space densities
## Abstract
The low velocity part of a kinetic equilibrium dark matter distribution has higher phase space density and is more easily incorporated in formation of a low mass galaxy than the high velocity part. For relativistically decoupling fermions (bosons), this explains one (two) orders of magnitude of the observed trend, that phase space densities in dark matter halo cores are highest in the smallest systems, and loosens constraints on particle masses significantly. For non-relativistic decoupling and/or finite chemical potentials even larger effects may occur. It is therefore premature to dismiss dissipationless particle distributions as dark matter on the basis of phase space arguments.
It has recently become clear, that the otherwise successful cold dark matter model (CDM) for cosmic structure formation has several severe problems, such as predictions of cusps in the central density profiles of galaxies, too many low-mass subclumbs within dark matter halos, and a lack of angular momentum in galaxy disks compared to observations. In one way or the other these problems are all related to the fact that CDM has no initial velocity spread, and therefore infinite initial density in phase space. Selfinteractions among the dark matter particles has been suggested as a possible solution. Another suggested solution has been the reintroduction of warm dark matter (WDM) consisting of particles with a moderate primordial velocity spread. Contrary to CDM thermal WDM particles lead to finite core density, but apparently require a high mass and therefore extremely early decoupling from the primordial plasma to account for the observed core phase space density of $`10^4M_{}/\mathrm{pc}^3(\mathrm{km}/\mathrm{s})^3`$ in dwarf spheroidal galaxies. There may also be difficulties explaining the decrease in phase space density by a factor 10โ100 for dwarf spirals and low surface brightness galaxies, and a further factor of 10โ100 for normal spirals. A single particle mass for WDM would appear to lead to a definite prediction for the central halo phase space density because of conservation of fine-grained phase space density (Liouvilles theorem).
However, whereas such arguments are correct in terms of the average phase space density, they do not take into account, that while originally almost uniform in real space, the fine-grained density is an (exponentially) varying function of position in momentum space. Low momentum particles are in denser parts of phase space than high momentum particles, and depending on the actual distribution function (fermion or boson, zero or nonzero chemical potential, relativistic or nonrelativistic decoupling), the densest part of the distribution may have phase space density significantly above average. Furthermore, halo formation should typically include particles from the low momentum end first. A system with low gravitational potential like a dwarf spheroidal will effectively probe only the densest part of the phase space distribution, whereas a large spiral galaxy probably contains something close to the average phase density (possibly diluted by mergers etc).
Other effects (baryons, mergers, phase space dilution during gravitational collapse, particle selfinteractions etc) can further enhance the diversity in observed phase space densities in galaxy cores, and as demonstrated in a merging hierarchy where larger systems are gradually formed by merging of smaller units can explain many of the features observed. But as demonstrated in the following, even in the absence of mergers, a significant spread in primordial phase space densities in dark matter cores is predicted, with core densities decreasing with increasing escape velocity of the system in question. The magnitude of the effect ranges from one order of magnitude for relativistically decoupling, non-degenerate fermions, to several orders of magnitude for non-relativistically decoupling bosons. Ad hoc additions to these simplest distribution functions would allow a further range. Thus, there may still be room for a single dissipationless elementary particle explanation of dark matter in dwarf spheroidal as well as dwarf spiral galaxies. Furthermore, the natural selection of the highest phase density particles in the smallest systems leads to a significant reduction in the minimum mass for the particle responsible, loosening the rather strong requirements on the epoch of dark matter decoupling in the early Universe.
An isotropic gas of particles in kinetic equilibrium has a spatial number density
$$n=(g/h^3)4\pi f(p)p^2๐p,$$
(1)
where $`h`$ is Planckโs constant, $`g`$ is the number of helicity states, and
$$f(p)=\left\{\mathrm{exp}\left[(E\mu )/kT\right]\pm 1\right\}^1$$
(2)
is a Fermi-Dirac ($`+`$) or Bose-Einstein ($``$) distribution with chemical potential $`\mu `$. Energy $`E`$ is related to momentum $`p`$ and particle mass $`m`$ via $`E^2=(pc)^2+m^2c^4`$, where $`c`$ is the speed of light.
A particle species that decouples from the remaining plasma in the early Universe at temperature $`T_D`$ redshifts its momenta in proportion to the expansion of the Universe, $`p=p_DR_D/R`$, where $`R(t)`$ is the cosmic scale factor, and its number density (particle number is conserved) evolves like $`nR^3`$. From this follows that the distribution function $`f`$ at a time after decoupling is related to the distribution at decoupling $`f_D`$ from Eq. (2) like
$$f(p)=f_D(pR/R_D).$$
(3)
$`f(p)`$ even keeps an equilibrium shape after decoupling in two regimes , namely ultrarelativistic decoupling ($`Epc`$, $`T=T_DR_D/R`$, $`\mu =\mu _DR_D/R`$), with
$$f_R(p)=\left\{\mathrm{exp}\left[(pc\mu )/kT\right]\pm 1\right\}^1,$$
(4)
and nonrelativistic decoupling ($`E\mu p^2/2m\mu _{\mathrm{kin}}`$, $`\mu _{\mathrm{kin}}\mu mc^2=\mu _{\mathrm{kin},\mathrm{D}}(R_D/R)^2`$, $`T=T_D(R_D/R)^2`$), so
$$f_N(p)=\left\{\mathrm{exp}\left[(p^2/(2m)\mu _{\mathrm{kin}})/kT\right]\pm 1\right\}^1.$$
(5)
The distribution of fine-grained phase space density is conserved in time (Liouvilles theorem). This fact is expressed by Eq. (3), and it means that the phase space distribution at decoupling can be directly related to measurements of dark matter phase densities today. Applied to conservation of the maximum phase space density this was the basis for the Tremaine-Gunn limit on dark matter fermion masses, later generalized to bosons (where no maximum exists) by means of the average phase space density.
Recently Hogan and Dalcanton reconsidered the issue of the primordial average phase space density compared with observations of dark matter phase space densities in halo cores. From the distribution function they calculated mass density $`\rho =mn`$, and pressure $`P=(g/h^3)p^2/(3E)fd^3p(g/h^3)p^4/(3mc^2)4\pi ๐p=mn<v^2>/(3c^2)`$, performing the calculation in the nonrelativistic regime, where $`Emc^2`$ and $`p=mv`$. This leads to the following expression for the phase space density, $`Q`$, defined by
$$Q\frac{\rho }{<v^2>^{3/2}}=\frac{m^4g}{h^3}4\pi \frac{[f(p)p^2๐p]^{5/2}}{[f(p)p^4๐p]^{3/2}}.$$
(6)
Hogan and Dalcanton uses this expression in the form (units with $`\mathrm{}=c=1`$)
$$Q_X=q_Xg_Xm_X^4$$
(7)
for particle type $`X`$, where $`q_X=0.0019625`$ for a relativistically decoupling, $`\mu =0`$ fermion, and $`q_X=0.036335`$ for a relativistically decoupling, $`T=0`$, $`\mu m`$ degenerate fermion, where the $`q`$-values come from taking the complete integrals over the distribution function.
For the remainder of this investigation we return to dimensional units and keep factors of $`\mathrm{}`$ and $`c`$. In these units
$$q_XQ_Xh^3m^4g^1=4\pi \frac{[f(p)p^2๐p]^{5/2}}{[f(p)p^4๐p]^{3/2}}.$$
(8)
For a degenerate fermion with relativistic decoupling (limit $`(m\mu )/kT_D\mathrm{}`$) this gives $`q_{RFdeg}=4\pi 5^{3/2}/3^{5/2}=9.0128415`$, so $`\overline{f}=q/9.0128415`$ is in general a measure of the average occupation number in a phase space distribution (in units of $`g/h^3`$). For a zero chemical potential fermion the corresponding average $`q`$-values for relativistic and non-relativistic decoupling (in the latter case zero chemical potential means $`\mu _{\mathrm{kin}}=0`$) are $`q_{RF0}=0.4868039`$ and $`q_{NF0}=1.9223`$, whereas the similar numbers for bosons are $`q_{RB0}=0.9071055`$ and $`q_{NB0}=21.521`$. Notice the larger values for bosons, that express the fact that bosons have a higher fraction of low momentum, high phase space density particles than fermions.
So far only average characteristics of the distributions have been discussed, but it turns out to be quite interesting to study the whole distribution of phase densities. Figure 1 shows these distributions in a plot of $`q(p)/q_X`$ as a function of the fraction of particles with momentum less than $`p`$ (calculated at decoupling, but the distribution is conserved in time by Liouvilles theorem). Here $`q_X`$ is given in Eq. (8) integrating from 0 to $`\mathrm{}`$, whereas $`q(p)`$ is defined by the same equation, but integrating only from 0 to $`p`$. The ratio therefore illustrates the amplification of phase space density relative to the average for a given dark matter distribution function if only the densest parts of phase space are utilized, for instance in formation of a galaxy halo.
Notice that relativistically decoupling fermions show an order of magnitude amplification, whereas two orders of magnitude can be gained for relativistically decoupling bosons and several orders of magnitude for nonrelativistically decoupling bosons.
Another way of illustrating the amplification effect is shown in Figure 2, where amplification is plotted as a function of the dimensionless momentum $`xp_Dc/kT_D`$, which is the natural integration variable in the calculations. For particles that are nonrelativistic at the epoch of galaxy formation, $`R_g`$, $`x`$ is related to particle speed $`v`$ at that time by
$$x=\frac{v}{c}\frac{mc^2}{kT_D}\frac{R_g}{R_D}.$$
(9)
The occurrence of a phase space amplification factor at low momenta is a natural consequence of Eq. (2). The fine-grained occupation number (and therefore also the coarse-grained phase space occupation) has a maximum at $`p=0`$ equal to $`f_{\mathrm{max}}=\left\{\mathrm{exp}\left[(mc^2\mu )/kT\right]\pm 1\right\}^1`$, which is 1 for degenerate fermions ($`(mc^2\mu _D)/kT_D\mathrm{}`$), 0.5 for fermions decoupling when $`(mc^2\mu _D)/kT_D=0`$, diverges for bosons in the same limit, and equals $`\mathrm{exp}\left[(\mu _Dmc^2)/kT_D\right]`$ for fermions and bosons in the limit $`(mc^2\mu _D)/kT_D\mathrm{}`$. Quantitatively, the amplification factor for a relativistically decoupling fermion behaves like $`9.26(15x/16)`$ to first order in $`x`$ (Fig. 2), or $`9.26(10.691F_F^{1/3}`$) expressed in terms of the fraction of fermions, $`F_F`$ (Fig. 1). The similar limits for bosons are $`19.59x^1(17x/30)`$, or $`8.93F_B^{1/2}4.57`$ (notice that these factors diverge for small $`x`$ or $`F_B`$).
Using entropy conservation in the cosmic expansion $`R_g`$ and $`R_D`$ entering the equation for $`x`$ are related by $`g_gT_{\gamma g}^3R_g^3=g_DT_{\gamma D}^3R_D^3`$, where $`g_{}`$ counts the total number of effective particle degrees of freedom at the given epoch. Today and at galaxy formation, $`g_g=43/11`$. Introducing the redshift of galaxy formation $`z_g`$ via $`T_{\gamma g}=(1+z_g)T_{\gamma 0}`$, where the present photon temperature is $`T_{\gamma 0}=2.726\mathrm{K}`$, $`x`$ can be expressed as
$$x=0.0223\left(\frac{mc^2}{1\mathrm{e}\mathrm{V}}\right)\left(\frac{10}{1+z_g}\right)\left(\frac{v}{10\mathrm{k}\mathrm{m}/\mathrm{s}}\right)g_D^{1/3}.$$
(10)
The value of $`g_D`$ depends on the epoch of decoupling. For standard neutrino decoupling at $`kT_D1\mathrm{M}\mathrm{e}\mathrm{V}`$, $`g_D=43/4`$. But much higher values of $`g_D`$ are possible for earlier decoupling, with $`g_D50`$ above the quark-hadron phase transition temperature, $`kT100\mathrm{M}\mathrm{e}\mathrm{V}`$, $`g_D100`$ above $`kT200\mathrm{G}\mathrm{e}\mathrm{V}`$, and even higher values possible earlier on.
The value of $`g_D`$ not only determines the $`x`$-$`v`$ relation, but also crucially impacts on the total mass density contribution of the particle in question, and thereby determines its potential as a dark matter candidate. For the case of relativistic decoupling ($`(m\mu _D)/T_D=0`$, $`m/T_D0`$) fermions contribute to the cosmic density like
$$\mathrm{\Omega }_Xh^2=0.0572\left(\frac{mc^2}{1\mathrm{e}\mathrm{V}}\right)g/g_D,$$
(11)
where $`h`$ is now the Hubble-parameter in units of 100 km s<sup>-1</sup> Mpc<sup>-1</sup>, and a similar expression (multiplied by $`4/3`$) applies for bosons.
It is interesting to note, that an explanation of the highest phase space densities measured, those in dwarf spheroidals of order $`10^4M_{}\mathrm{pc}^3(\mathrm{km}/\mathrm{s})^3`$, for a $`g=1`$ boson requires a particle mass of only 224 eV (309 eV) if the densest 1% (10%) of the bosons are used, increasing to 681 eV for the average occupation number (masses are reduced by a factor $`2^{1/4}`$ if $`g=2`$). Such masses are not in conflict with estimates of cosmic density for reasonably high values of $`g_D`$, and $`x`$ can indeed be low enough for reasonable values of $`z_g`$ to select only the densest part of phase space for dwarf spheroidals with typical velocity dispersions below 10 km/s.
For fermions with $`g=2`$ the high-phase space density selection expected for formation of dwarf spheroidals reduces warm dark matter particle mass limits from 669 eV to 383 eV, again loosening constraints on $`g_D`$ .
Even stronger effects may occur in a non-relativistic decoupling regime (upper dotted curve in Fig. 1), or one might consider adding extra features to the dark matter distribution functions, such as a small amount of Bose-Einstein condensation in the zero momentum, infinite phase density part of a boson distribution.
Such โfine-tuningโ may further loosen constraints on dissipationless dark matter, but even without it, part of the observed trend for the core phase space density to decrease with increasing gravitational potential when going from dwarf spheroidal to dwarf spiral galaxies is naturally explained by the selection of low-momentum dark matter particles described here. At the same time constraints on WDM particle masses and decoupling epochs are significantly reduced. These considerations should be taken into account and tested in detailed numerical simulations of dark matter halo formation, which are needed to settle the question of whether dissipationless particles may after all account for the dark matter in galaxies.
This work was supported in part by the Theoretical Astrophysics Center under the Danish National Research Foundation.
|
warning/0006/cond-mat0006177.html
|
ar5iv
|
text
|
# On the interpretation of spin-polarized electron energy loss spectra
## I Introduction
The study of elementary excitations in itinerant-electron ferromagnets is an area which is currently very active in spite of the enormous amount of publications on the subject since early work in the sixties (see e.g. Refs. and references therein). From the beginning, experimental and theoretical work on these materials concentrated on neutron scattering and dynamical susceptibility studies. Efforts have continued in this direction until today, both because of the gradual improvement of electronic band structure calculations and because of the improvement of the experimental method. Around the mid-eighties, however, a new technique was introduced in the field, namely, spin polarized electron energy loss spectroscopy (SPEELS). Among its first successes, one can count the first observations, in a ferromagnetic glass and in nickel, of what were interpreted as Stoner excitations. Further work, reporting more detailed measurements, confirmed those findings. Theoretical model calculations of inelastic electron spin-flip exchange scattering, provided a basis for the interpretation of those observations in terms of Stoner excitations. In addition, Vignale and Singwi found in their work that spin waves should also be observable in SPEELS measurements. However, these had not been observed at the time, nor were they observed in the several years that followed. Spin waves were found in other model calculations, the calculations of Plihal and Mills being the most conclusive in this respect because of their more accurate treatment of electronic structure. It is only very recently that the detection of spin waves in a SPEELS experiment has finally been reported. The application of SPEELS has been naturally extended to the study of magnetic surfaces. An important theoretical effort in this direction is that by Mills and collaborators, who have studied ferromagnetic thin films as well.
To introduce the questions addressed by this work, we recall briefly some of the main concepts involved in SPEELS and discuss some of the findings to date. SPEELS is a spin-polarized version of electron energy loss spectroscopy in the sense that the spin polarization of the scattered electrons is also measured. The impinging electron often is also spin-polarized, but this is not necessarily so (see e.g. Refs. ). In a so-called spin-flip exchange scattering event, an incoming electron with given spin comes to occupy an empty level in the material while an electron with opposite spin is driven out and is detected. The process produces thus a Stoner excitation. In the band picture of magnetic transition metals, below the Curie temperature, the exchange split 3$`d`$ bands provide large densities of occupied majority-spin states below the Fermi energy and vacant minority-spin states above it. Thus, it is more likely for an impinging electron with minority spin to excite a Stoner pair than for a majority-spin incoming electron, particularly for an excitation energy corresponding to the exchange splitting of the ferromagnet. This is the mechanism invoked to explain the Stoner peak or the asymmetry reported in Refs. and and further experimental work (Refs. ). However, it turned out necessary to elaborate on several other issues. Firstly, the Stoner peak was very broad in all observations. This was interpreted by Kirschner, Rebenstorff, and Ibach as an indication of the nonuniformity of exchange splitting throughout the Brillouin zone. Then, in Fe, the energy loss at which the Stoner peak occurs and its width were reported by Venus and Kirschner to increase with increasing scattering angle, a fact that was correlated by these authors with the calculated density of Stoner states. Also, a threshold for the onset of Stoner excitations in Ni(110) was reported by Abraham and Hopster and interpreted in terms of the Ni 3$`d`$ band structure. These workers, moreover, indicated that their spectra did not differ significantly for off specular scattering angles ranging from 10 to 40, which they explained as due to the non-conservation of the momentum component perpendicular to the surface.
Finally, an important application based on SPEELS interpretation is that, in surface and thin film studies, the Stoner peak is assumed to give information on the surface magnetic moment through the correlation between exchange splitting and moment. In particular, a Stoner peak found at higher energies than the exchange splitting bulk value is assumed to indicate an enhanced magnetic moment at the surface.
Clearly, more theoretical work is required, for the bulk as much as for surfaces, to make further progress. In particular, it would be important to understand better the phenomenology of SPEELS and to try to be more specific about the information we can expect from it. This would also provide experimenters with useful feedback. Accordingly, we think it is worthwhile going back to a model calculation and look more closely at the dynamic properties of the material probed by SPEELS. In this work we consider a model of Fe based on paramagnetic tight-binding 3$`d`$ bands, with up and down spin bands rigidly split. The cross section for spin-flip exchange scattering processes is evaluated within the random phase approximation, assuming the solid is described by a multi-band Hubbard Hamiltonian. As we shall see, this allows us to show that it is not the total density of Stoner states as a function of energy loss and momentum transfer which causes the structure in the Stoner region of the spectrum, but the density of Stoner states for a few interband excitations. Which interband excitations are important is essentially determined by the weight of the matrix elements for such processes. It this regard the contribution of umklapp scattering is fundamental because of the the coupling of different bands at different energy ranges. This gives rise to a richer structure in the Stoner region of the spectra. Also, our model predicts that optical spin waves should be observable through SPEELS. Again umklapp scattering proves critical, providing optical spin waves with the necessary oscillator strength. We find this strength to depend importantly on scattering angle.
Section II of this paper is devoted to theory, presenting the derivation of the spin-flip exchange scattering cross section for our model. We present our main results in Section III. Then follows in Section IV a discussion of our results in the light of experimental findings and other theoretical work. Finally, in Section V we summarize our work and give some conclusions.
## II Theory
### A Spin-flip exchange scattering cross section
The electron spin-flip exchange scattering differential cross section for a $`N`$-electron system target has been previously derived on general grounds by Vignale and Singwi in terms of a particle-hole excitation correlation function. One has
$$\frac{d^2\sigma }{dEd\mathrm{\Omega }}=\frac{m^2}{4\pi ^2\mathrm{}^4}\frac{p_f}{p_i}\frac{1}{\pi }\frac{\mathrm{Im}\chi _{\sigma _i\sigma _f}^R(๐ฉ_i,๐ช,E)}{1e^{\beta E}},$$
(1)
where $`E`$ is the energy loss, $`\mathrm{\Omega }`$ is the solid angle, $`m`$ the electron mass, and $`\beta =1/k_BT`$. Momentum transfer is given by $`๐ช=๐ฉ_i๐ฉ_f`$, with $`๐ฉ_i`$ and $`๐ฉ_f`$ the momentum of the incoming and outgoing electrons, respectively. Likewise, $`\sigma _i`$ is the spin of the impinging electron and $`\sigma _f`$ that of the scattered one. The retarded function $`\chi _{\sigma _i\sigma _f}^R`$ can be obtained by analytic continuation of the two-particle temperature correlation function
$`\chi _{\sigma _i\sigma _f}(๐ฉ_i,๐ช,i\omega _n)`$ $`={\displaystyle _0^\beta }๐\tau e^{i\omega _n\tau }`$ (3)
$`\times T_\tau [\varrho _{\sigma _i\sigma _f}(๐ฉ_i,๐ช,\tau )\varrho _{\sigma _i\sigma _f}^{}(๐ฉ_i,๐ช)].`$
In this equation $`\omega _n=2\pi n/\beta `$ is a bosonic Matsubara frequency, $`T_\tau `$ is the imaginary time ordering operator, and the brackets indicate the thermodynamic average in the canonical ensemble. $`\varrho _{\sigma _i\sigma _f}^{}`$ is the particle-hole creation operator
$`\varrho _{\sigma _i\sigma _f}^{}(๐ฉ_i,๐ช)={\displaystyle \frac{1}{N}}`$ $`{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle ๐๐ซ๐๐ซ_je^{i๐ฉ_f๐ซ_j}e^{i๐ฉ_i๐ซ}}`$ (5)
$`\times v(|๐ซ๐ซ_j|)\psi _{\sigma _i}^{}(๐ซ)\psi _{\sigma _f}(๐ซ_j),`$
where $`\psi _\sigma ^{}(๐ซ)`$ is the field operator creating an electron of spin $`\sigma `$ at position $`๐ซ`$ and $`v(r)=e^2/r`$ is the Coulomb interaction between the scattered and target electrons. The sum runs over the $`N`$ electrons in the target system. This expression is quite general, and could be applied equally well to a solid, an atom, or a molecule.
We now consider the $`N`$ electrons in a crystal material. We write the Bloch wave function for a state with wave vector $`๐ค`$ and spin $`\sigma `$ in band $`n`$ in terms of Wannier functions,
$$\psi _{n๐ค\sigma }(๐ซ)=\frac{1}{\sqrt{N_0}}\underset{๐}{}e^{i๐ค๐}\varphi _{n๐ค}(๐ซ๐)\eta _\sigma ,$$
(6)
where $`N_0`$ is the number of sites in the crystal and $`\eta _\sigma `$ is the spin function. Denoting by $`a_{n๐ค\sigma }^{}`$ the operator creating an electron in such a state, the field operators can be expanded as $`\psi _\sigma ^{}(๐ซ)=_{n๐ค}\psi _{n๐ค\sigma }(๐ซ)a_{n๐ค\sigma }^{}`$. The particle-hole creation operator becomes
$$\varrho _{\sigma _i\sigma _f}^{}(๐ฉ_i,๐ช)=\underset{nn^{}}{}\underset{๐ค}{}W_{nn^{}}(๐ฉ_i,๐ช,๐ค)a_{n๐ค\sigma _i}^{}a_{n^{}๐ค๐ช\sigma _f}^{},$$
(7)
where the sum in momentum space runs over the Brillouin zone and matrix element $`W_{nn^{}}`$ is given by
$`W_{nn^{}}(๐ฉ_i,๐ช,๐ค)={\displaystyle \frac{N_0}{V}}{\displaystyle \underset{๐}{}}\widehat{v}`$ $`(๐ค๐ฉ_i๐)\widehat{\varphi }_{n๐ค}^{}(๐ค๐)`$ (9)
$`\times \widehat{\varphi }_{n^{}๐ค๐ช}(๐ค๐ช๐).`$
Here $`๐`$ denotes vectors in the reciprocal lattice and $`V`$ is the volume of the sample. The $`\widehat{}`$ indicates a Fourier transformed function and denotes complex conjugation. To write the last two equations we have defined $`a_{n๐ค+๐\sigma }^{}a_{n๐ค\sigma }^{}`$ and have exploited the periodicity of the Wannier functions in the wave vector index, i.e. $`\varphi _{n๐ค+๐\sigma }=\varphi _{n๐ค\sigma }`$.
### B RPA expression for a tight-binding system
We are interested in the cross section for an itinerant electron ferromagnet. We describe the system within a tight-binding approximation, thus writing the Wannier wave function for given $`๐ค`$ and band index $`n`$ by a linear combination of atomic orbitals $`\phi _m`$
$$\varphi _{n๐ค}(๐ซ)=\underset{m}{}b_{mn}(๐ค)\phi _m(๐ซ).$$
(10)
The coefficients $`b_{mn}`$ diagonalize the crystal Hamiltonian and are normalized so as to define a unitary matrix. Consequently, the independent-electron Hamiltonian of the system can be written $`H_0=_{n๐ฉ\sigma }ฯต_n(๐ฉ)a_{n๐ฉ\sigma }^{}a_{n๐ฉ\sigma }^{}`$, where the $`ฯต_n(๐ฉ)`$ are paramagnetic band energies. We assume the interacting system is described by a multi-band Hubbard Hamiltonian. In our basis we have
$`H_I={\displaystyle \frac{1}{2}}{\displaystyle \frac{U}{N_0}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{nn^{}}{mm^{}}}{}}`$ $`{\displaystyle \underset{\sigma }{}}{\displaystyle \underset{\mathrm{๐ฉ๐ฉ}^{}๐ช^{}}{}}c_{nm}(๐ฉ+๐ช^{},๐ฉ)c_{n^{}m^{}}(๐ฉ^{}๐ช^{},๐ฉ^{})`$ (12)
$`\times a_{n๐ฉ+๐ช^{}\sigma }^{}a_{m๐ฉ\sigma }a_{n^{}๐ฉ^{}๐ช^{}\sigma }^{}a_{m^{}๐ฉ^{}\sigma }^{},`$
where we have defined $`c_{nm}(๐ฉ,๐ช)=_lb_{ln}(๐ฉ)b_{lm}(๐ช),`$ and $`U`$ is the effective on-site Coulomb interaction for two electrons with opposite spin. The RPA evaluation of the correlation function $`\chi _{\sigma _i\sigma _f}`$ defined in Eq. 3 is a straightforward generalization of that in previous work . The response function divides naturally in two,
$$\chi _{\sigma _i\sigma _f}(๐ฉ_i,๐ช,i\omega _n)=\chi _{\sigma _i\sigma _f}^\mathrm{S}(๐ฉ_i,๐ช,i\omega _n)+\chi _{\sigma _i\sigma _f}^{\mathrm{MB}}(๐ฉ_i,๐ช,i\omega _n).$$
(13)
The Stoner or single-particle excitation contribution is given by
$`\chi _{\sigma _i\sigma _f}^\mathrm{S}(๐ฉ_i,๐ช,i\omega _n)={\displaystyle \underset{nn^{}}{}}{\displaystyle \underset{๐ค}{}}`$ $`{\displaystyle \frac{f_{n^{}๐ค๐ช\sigma _f}f_{n๐ค\sigma _i}}{i\omega _n+ฯต_{n^{}\sigma _f}(๐ค๐ช)ฯต_{n\sigma _i}(๐ค)}}`$ (15)
$`\times |W_{nn^{}}(๐ฉ_i,๐ช,๐ค)|^2.`$
We have introduced the occupation probability of state $`n๐ค\sigma `$, $`f_{n๐ค\sigma }=a_{n๐ค\sigma }^{}a_{n๐ค\sigma }`$, and the single-particle energy modified by the exchange self-energy
$$ฯต_{n\sigma }(๐ค)=ฯต_n(๐ค)\frac{U}{N_0}\underset{๐ฉ}{}f_{n๐ฉ\sigma }.$$
(16)
Thus, in this model, spin down and spin up energy bands are rigidly split by the quantity $`\mathrm{\Delta }=U(n_{}n_{}),`$ where
$$n_\sigma =\frac{1}{N_0}\underset{m๐ฉ}{}f_{m๐ฉ\sigma }$$
(17)
is the average number per site of states with spin $`\sigma `$.
The many-body contribution is given by
$`\chi _{\sigma _i\sigma _f}^{\mathrm{MB}}(๐ฉ_i,๐ช,i\omega _n)={\displaystyle \frac{U}{N_0}}{\displaystyle \underset{nn^{}}{}}`$ $`G_{\sigma _i\sigma _f}^{nn^{}}(๐ฉ_i,๐ช,i\omega _n)`$ (19)
$`\times \mathrm{\Gamma }_{\sigma _i\sigma _f}^{nn^{}}(๐ฉ_i,๐ช,i\omega _n),`$
with the auxiliary functions $`G^{nn^{}}`$ and $`\mathrm{\Gamma }^{nn^{}}`$ defined as follows:
$`G_{\sigma _i\sigma _f}^{nn^{}}(๐ฉ_i,๐ช,`$ $`i\omega _n)={\displaystyle }_{mm^{}}{\displaystyle }_๐ค{\displaystyle \frac{f_{m^{}๐ค๐ช\sigma _f}f_{m๐ค\sigma _i}}{i\omega _n+ฯต_{m^{}\sigma _f}(๐ค๐ช)ฯต_{m\sigma _i}(๐ค)}}`$ (21)
$`\times W_{mm^{}}^{}(๐ฉ_i,๐ช,๐ค)b_{nm}(๐ค)b_{n^{}m^{}}(๐ค๐ช),`$
and, considering $`G`$ and $`\mathrm{\Gamma }`$ as vectors with coefficients indexed by $`nn^{}`$,
$`\mathrm{\Gamma }_{\sigma _i\sigma _f}(๐ฉ_i,๐ช,i\omega _n)=`$ $`\left[1+{\displaystyle \frac{U}{N_0}}D_{\sigma _i\sigma _f}(๐ช,i\omega _n)\right]^1`$ (23)
$`\times G_{\sigma _i\sigma _f}(๐ฉ_i,๐ช,i\omega _n),`$
where the elements of matrix $`D`$ are
$`D_{\sigma _i\sigma _f}^{nn^{},mm^{}}(๐ช,i\omega _n)={\displaystyle \underset{ll^{}}{}}{\displaystyle \underset{๐ค}{}}b_{nl}(๐ค)b_{n^{}l^{}}(๐ค๐ช)`$ (24)
$`\times {\displaystyle \frac{f_{l^{}๐ค๐ช\sigma _f}f_{l๐ค\sigma _i}}{i\omega _n+ฯต_{l^{}\sigma _f}(๐ค๐ช)ฯต_{l\sigma _i}(๐ค)}}b_{ml}(๐ค)b_{m^{}l^{}}(๐ค๐ช).`$ (25)
To obtain $`\chi _{\sigma _i\sigma _f}^R`$, analytic continuation $`i\omega _nE+i\eta `$ of the above results is straightforward. Also, results in next section are obtained in the zero-temperature limit.
## III Results
Before presenting our results, there are a few details of our model we should discuss. As we said in the introduction, our Fe model is based on simple $`3d`$ tight-binding paramagnetic bands, i.e. we neglect $`sp`$ hybridization. This has, of course, incidence on the quantitative details of the response of the system. However, we will see our results are rather general and do not depend on these details. The Wannier functions were written as a linear combination of the Fe five 3$`d`$ atomic wave functions. The overlap integrals were calculated, up to next nearest neighbors, using the Fe atomic wave functions determined according to Griffithโs prescription, and a lattice constant $`a=2.87`$ ร
. We have, thus, a five band model. The bandwidth was set to 4.7 eV, which corresponds roughly to the bandwidth of $`d`$ electrons in Fe. Exchange splitting $`\mathrm{\Delta }`$ was chosen to be 2 eV, taking as reference the position of the peaks in the densities of states for up and down spins. Then, the Fermi level was fixed by the condition of having six electrons per unit cell. We show the density of states for up and down spins in Fig. 1. These exhibit the bonding and antibonding regions common to BCC materials with unfilled $`d`$ shells. Once the Fermi energy is fixed we can deduce the strength of the effective Coulomb interaction $`U`$ in our model from $`U=\mathrm{\Delta }/(n_{}n_{}).`$ We find $`U=0.69`$ eV, which compares very well with the energy found in other works. On the other hand, bulk polarization is too high, roughly 48%, reflecting the lack of hybridization with $`sp`$ electrons.
In a typical SPEELS experiment, the incoming electron beam impinges on the sample surface at an angle $`\theta `$ to the normal and the total scattering angle is 90. For fixed scattering angle, i.e. for given incoming and outgoing momenta, there are three possible scattering processes corresponding to different momentum transfer. In two cases, a relatively small angle inelastic scattering event is preceded or followed by elastic scattering. In the third one, all the momentum transfer is absorbed by the electron-hole pair excitations, i.e. it is a large angle scattering event. We consider here the geometry of Venus and Kirschner, that is, the sample exposes the (110) surface and the scattering plane is defined by the surface normal and the axis. If $`u`$ is the axis normal to the surface, the three momenta mentioned are given as a function of energy loss $`E`$ and impinging momentum $`p_i`$ and energy $`E_i`$ by
$`q_u=p_i(\mathrm{cos}\theta \mathrm{sin}\theta \sqrt{1E/E_i}),`$ (26)
$`q_u^{}=p_i(\mathrm{cos}\theta +\mathrm{sin}\theta \sqrt{1E/E_i}),`$ (27)
$`q_u^{\prime \prime }=p_i(\mathrm{cos}\theta +\mathrm{sin}\theta \sqrt{1E/E_i}),`$ (28)
$`q_z=q_z^{}=q_z^{\prime \prime }=p_i(\mathrm{sin}\theta \mathrm{cos}\theta \sqrt{1E/E_i}),`$ (29)
where $`\theta `$ is the angle to the normal. The large angle scattering event corresponds to $`๐ช^{\prime \prime }`$. Momentum transfer parallel to the surface is the same in the three cases. To calculate the final spectrum, the contributions of these three processes have to be added because experiment does not discriminate them.
Also, since Fe presents a non negligible quantity of free-like $`s`$ and $`p`$ states at the Fermi surface, the interaction between the incoming electrons and those in the solid will be screened. We take this into account using the Thomas-Fermi form of the screened Coulomb potential, with a screening wave vector corresponding to the density of states of $`s`$ and $`p`$ electrons at the Fermi surface. This gives $`q_{\mathrm{TF}}=0.26`$ in units of $`k_a=4\pi /a`$. Finally, an important point to mention is that we use a finite value for $`\eta `$ when taking the analytic continuation $`i\omega _nE+i\eta `$. Since, to our knowledge, there are no estimates of the self-energy corrections for up and down spin bands in Fe, we take $`\eta =80`$ meV, which corresponds to the resolution in the latest experiment on this material.
### A Interband densities of Stoner states
Let us consider a majority spin electron with an angle of incidence $`\theta =60^{}`$ to the normal and an incoming energy $`E_i=22`$ eV, which is the energy used in Ref. . We see in Fig. 2(a) that the total spin-flip exchange scattering cross section is indeed rather broad, with its peak centered at an energy much higher than the exchange splitting value (2 eV in our model), a trend observed experimentally by Venus and Kirschner. We also show the partial cross sections for different momentum transfers. The curves for small scattering angle coincide for
symmetry reasons. Though total cross section broadness is somewhat increased because of the difference between small angle and large angle scattering, the cross section in each case is broad in itself. We have examined the origin of the structure in this spectrum. Firstly, we separated single-particle excitations and many-body effects. In Fig. 2(b) we show the noninteracting and interacting (total) cross sections. This figure clearly shows us the contributions of collective modes. Indeed, the broad feature starting around 0.2 eV indicates the excitation of low lying spin waves, and, more interestingly, the shoulder at higher energy, below 2 eV, indicates the excitation of optical spin waves. This is important because optical spin waves have not been discussed previously in connection with SPEELS measurements. We consider spin waves again further on and we concentrate here on the single particle traits. Besides the peak at 3 eV, the noninteracting cross section shows a shoulder at 4 eV and a broad feature, albeit much smaller, at low energy loss, around 1 eV or so. SPEELS spectra have often been interpreted in terms of the density of Stoner states. Accordingly, we show in Fig. 2(c) the total density of Stoner states, as well as the densities of Stoner states for different momentum transfer (as before, the curves for small angle scattering are the same). It is evident in the cross sections that there is nothing reminiscent of the high density of states at the exchange splitting energy. The only features of the cross sections that can find an explanation in the density of Stoner states are the shoulder at 4 eV and, possibly, the hump around 1 eV. We have, thus, refined our study and have considered the behavior of the density of Stoner states as a function of energy loss and the bands coupled in an excitation (recall energy loss and momentum transfer are coupled, cf. Eq. (29))
$`\rho _{nn^{}}(E)={\displaystyle \frac{1}{N_0}}`$ $`{\displaystyle \underset{๐ค}{}}(f_{n^{}๐ค๐ช\sigma _f}f_{n๐ค\sigma _i})`$ (31)
$`\times \delta (E+ฯต_{n^{}\sigma _f}(๐ค๐ช)ฯต_{n\sigma _i}(๐ค)).`$
Hence, subscripts $`n`$ and $`n^{}`$ indicate minority and majority bands, respectively (bands are numbered from bottom to top). We show a plot of the density of states thus defined in Fig. 3(a). We can see that the allowed and forbidden interband excitations are completely identified (we use the term forbidden for interband excitations with vanishing density of Stoner states). Moreover, the series of Stoner peaks clearly reflect the bonding and antibonding nature of the electronic structure, giving rise to two arrays of peaks, for higher and lower excitation energies. The question is, of course, which of these Stoner peaks contribute the most to SPEELS cross sections. The answer is to be found, perhaps unsurprisingly, in how strongly the different bands are coupled by the matrix elements $`W_{nn^{}}`$ of the electron-hole creation operator, which weights the contribution of each Stoner excitation (see Eq. 15). What is not so obvious is the outcome of the combined effect of Stoner peaks and matrix elements. Let us consider the average of the square of the absolute value of matrix elements $`W_{nn^{}}`$ over the Brillouin zone, i.e.
$$|W_{nn^{}}|^2=\frac{1}{\widehat{v}}\underset{๐ค}{}|W_{nn^{}}|^2$$
(32)
($`\widehat{v}`$ denoting the volume of the Brillouin zone). In Fig. 3(b) we show the graph of $`|W_{nn^{}}|^2`$ as a function of energy loss and of bands coupled. There is little significant variation as a function of energy loss, but a very important structure as a function of band couple, resulting in a wave like pattern. Comparing Figs. 3(a) and 3(b) we can clearly see when it is that both quantities, $`\rho _{nn^{}}`$ and $`|W_{nn^{}}|^2`$, interfere constructively. Thus, although the density of Stoner states reaches is highest peak at exchange splitting, the average $`|W_{nn^{}}|^2`$ is negligible for the corresponding band couples. Instead, although the densities of Stoner states for interband excitations 21 and 22, 31 and 32, and 41 and 42 are more modest, the corresponding matrix element averages are high. Whence the peak around 3 eV and the shoulder around 4 eV in the noninteracting cross section in Fig. 2(b). Actually, the peak at 3 eV is more of a hat on top of the high cross section value due to interband excitations 31 and 32. We can also see that the hump around 1 eV is due to excitations coupling bands 2 and 3, and 2 and 4. To corroborate our analysis, we show in Fig. 4 the cross section taking into account solely the interband processes mentioned above. We include the total noninteracting cross section for comparison as well. We see that the few interband excitations considered indeed account almost completely for the structure of the noninteracting spectrum. An argument to understand how so simple a picture can work is that, since the atomic 3$`d`$ orbitals are localized, their Fourier transform is rather flat, so that $`|W_{nn^{}}|^2`$ in the single-particle correlation function $`\chi _{\sigma _i\sigma _f}^\mathrm{S}`$ (cf. Eq. (15)) may be replaced by its average value over the Brillouin zone. Thus, $`\chi _{\sigma _i\sigma _f}^\mathrm{S}`$ is approximately proportional to $`_{nn^{}}\rho _{nn^{}}|W_{nn^{}}|^2`$, a weighted average of the interband densities of Stoner states.
### B Umklapp processes
Another most interesting phenomenon playing a fundamental role in SPEELS is umklapp scattering. One can see in Eq. (9) that the contribution of umklapp processes to the particle-hole excitation operator $`\varrho _{\sigma _i\sigma _f}`$ is weighted by the Coulomb interaction and the Wannier wave functions. Because of the decay of the Coulomb potential as well as of the atomic orbitals with increasing wave vector, the weight becomes rapidly negligible for lattice vectors beyond first nearest neighbors. This is enough, however, for umklapp processes to have a twofold effect. To see this, let us consider cross sections taking into account only normal excitations (i.e. with respect to the first Brillouin zone). We show this in Fig. 5, where we plot both the interacting and noninteracting cross sections. Firstly, we see that, quite apart from their much lower values in comparison with the full response case, spectra in Fig. 5 show little resemblance with those in Fig. 2(b) (scale in both figures is the same). This is because the possible interband excitations have been drastically reduced. Indeed, let us consider the graph of the average $`|W_{nn^{}}|^2`$ for excitations strictly conserving crystal momentum. We show this in Fig. 3(c). The wave crests have been reduced to that for $`nn^{}=31,32`$, from which it is obvious the the different wave crests in Fig. 3(b) are due to umklapp scattering. Normal scattering alone results in a spectrum almost completely distorted because of the excitation of interband transitions mainly for energies between 3 and 4 eV (cf. Fig. 3(a)).
Also, in Fig. 5 it is immediately apparent that there remains no trace of spin waves in the spectrum. Indeed, the interacting and noninteracting curves are almost indistinguishable, with no hint of the spin wave modes below 2 or 1 eV. Thus, it appears that umklapp processes provide collective excitations with oscillator strength, at
least optical modes. We must point out here that the low lying collective mode contributing to the broad feature below 1 eV is not an acoustic mode, but also an optical one. We have calculated its dispersion relation in the direction and found that it tends linearly to 218.7 meV for $`q0`$. The slope is positive, but very low, with a value of 8.9 meV$``$ร
. The reason is that the energy bands in our model are purely $`d`$. It is well known that models of itinerant ferromagnetism which do not take into account hydridization with $`sp`$ bands, fail to describe spin waves appropriately. So we do not expect our model to predict accurate dispersion relations for collective excitations. However, we do think our result properly introduces optical spin waves as a source of structure in SPEELS measurements.
### C Spectra and angle of incidence
To illustrate further the pertinence of our analysis, let us consider an example. A question addressed in the past by experimenters, without finding a clear answer, has been that of the variation of spectra with scattering angle. Since $`|W_{nn^{}}|^2`$ plays such a consequential role, we take a look at its dependence on angle of incidence through the contour plot in Fig. 6. Generally speaking, the same interband averages remain the most important important as angle changes, except toward specular scattering, when weights raise and shift significantly (then the highest values are to be found for $`nn^{}=22,33`$, and 44). Let us consider the spectra for angles of incidence $`\theta =55^{}`$ and $`\theta =70^{}`$. In the first case, Fig. 6 shows us that the importance of interband excitations $`nn^{}=41,42`$ is diminished with respect to the $`\theta =60^{}`$ case. For $`\theta =70^{}`$, it is band couples $`nn^{}=21,22`$ that are diminished. In this way, in fact, we hinder the contribution of umklapp processes coupling different bands. Regarding the densities of Stoner states, on the other hand, we found that changes from one angle to another are rather small, affecting essentially only the height of the peaks. This is because a change in angle will not change the energies at which there can be a Stoner excitation, but basically the number of these. Thus, considering the averages $`|W_{nn^{}}|^2`$, the scattering cross section for $`55^{}`$ should present almost no shoulder around 4 eV, since interband excitations 41 and 42 are weak for that angle. Likewise, we expect the scattering cross section for $`70^{}`$ to have a weaker peak around 3 eV, because of the negligible contribution of excitations for $`nn^{}=21,22`$. We can appreciate these effects in Fig. 7. Thus, given the peaks of the interband densities of Stoner states, the shift in the Stoner excitation maximum with varying scattering angle obeys to which interband excitations receive more weight, according to the corresponding matrix elements averages. This, in turn, depends on which umklapp processes gain more importance. Fig. 7 also shows that the maximum shift causes the broadness of the spectrum to increase with scattering angle.
Turning to spin waves, it is interesting to note that the strength of the higher optical mode decreases with scattering angle. Thus it appears that umklapp scattering is unable to transmit to it sufficient oscillator strength at higher scattering angles. On the other hand, through comparison with the noninteracting cross sections we found that the lower lying spin wave seems less affected by scattering angle.
## IV Discussion
We wish to discuss some of the issues considered in this work pertaining to other theoretical and experimental findings. First of all, as we have shown (cf. Fig. 2(b)), the maximum in the SPEELS spectrum does not necessarily correspond to the exchange splitting energy of the ferromagnet, since the peak is at 3 eV and $`\mathrm{\Delta }=2`$ eV in our model. Thus, the common assumption that this peak allows us to estimate the magnetic moment through its correlation with exchange splitting should be reconsidered, both in bulk and surface studies.
Also, the broadness of the spectrum is generally associated with a non constant exchange splitting over the Brillouin zone. While we agree a non constant exchange splitting will have this effect, we have seen that a most important source of broadness is umklapp scattering, together with the structure of the interband densities of Stoner states of the material. In particular, even for specular or near specular scattering, will a model with rigidly split bands present a relatively broad spectrum. Still, according to our interband densities of Stoner states plot, for rigidly split bands we expect to see a maximum at exchange splitting. Unfortunately there are no reliable results for specular scattering in the case of Fe. In this regard, the case of Ni would be very interesting to investigate in more detail. To begin with, the results of Kirschner, Rebenstorff, and Ibach and of Abraham and Hopster are contradictory. Indeed, the former reported a broad maximum around exchange splitting for near specular scattering, while the latter stated that they see no sharp feature at that energy. According to our picture, Abraham and Hopsterโs result would be explained only if the matrix elements $`W_{nn^{}}`$ are always weak for energies near exchange splitting. An accurate calculation of the interband densities of Stoner states and of the weights $`W_{nn^{}}`$ would be most clarifying in this respect. For off specular scattering, however, both groups reported a weak dependence on scattering angle. This is plausible, according to our results. In Fig. 6 we see that the $`|W_{nn^{}}|^2`$ may remain relatively unchanged for certain scattering angle intervals, with, consequently, little change in the SPEELS spectra.
As mentioned in the introduction, moreover, Abraham and Hopster interpret the onset of Stoner excitations found in their work in terms of the 3$`d`$ band structure of Ni. Again, this could readily be verified having at hand the interband densities of Stoner states for this material. Indeed, these authors consider in particular interband excitations corresponding to $`nn^{}=55`$, the onset of which, if they exist, could be easily identified in a graph like that in Fig. 3(a). A point still to be verified would be if the matrix elements for such excitations are sufficiently important.
Spin waves, both acoustic and optical, have long been predicted in itinerant ferromagnets and subsequently observed through neutron scattering. The question is, then, if these are observable in SPEELS. The question to be studied in the future is if there can be enough coupling between the incoming electron and those in the solid to excite an optical spin wave. In our model, it is umklapp processes that provide optical spin waves with the necessary oscillator strength to contribute significantly to the spectra. It could be, however, the acoustic waves, when present, drain most of the oscillator strength. This is plausible because it is known that acoustic spin waves in Fe arise upon hybridization of $`d`$ electrons with $`sp`$ electrons. Matrix elements with $`sp`$ hybridization will be more important because of the larger $`s`$ wave functions. This implies, of course, as our model does, that optical modes are mainly $`d`$ in character. The case of Ni appears again to be different, since even a pure $`d`$ band model of Ni shows acoustic spin waves.
The analysis presented in this work can prove useful more broadly in the understanding of ferromagnetism. Recently, Hirsch has presented a model of ferromagnetism without exchange splitting, in which spin polarization arises upon broadening of the spin-up bands relative to spin-down. If this mechanism plays an important role in itinerant ferromagnets, then the interband densities of Stoner states will change considerably because other pairs of bands will be involved than those in the Stoner picture of ferromagnetism. Consequently, the predicted exchange scattering spectra will be different in both pictures. Thus SPEELS can prove a useful tool to validate or disprove Hirschโs model, possibly improving resolution previously.
Finally, we would like to comment on the bulk vs. surface question. Our calculations here have focused on bulk properties. However, some authors have presented SPEELS as a technique more appropriate for surface studies. The reason is concern regarding the mean free path of electrons at the energies used in SPEELS. The question raised, however, is not simple and requires more detailed consideration, both theoretically and experimentally. Still, most of the observations to date have been discussed in the light of bulk calculations. In this regard, a recent report on the electron dynamics at the surfaces of noble metals (Ag and Cu) is appropriate to mention. Bรผrgi et al. have found that the dynamics of hot electrons at surfaces can be dominated by bulk electrons. This offers more support to the premise that our results offer a sensible explanation for SPEELS results.
## V Summary and conclusions
In this work we address the problem of the interpretation of the SPEELS spectrum of itinerant ferromagnets. We find that considerably more information can be drawn from these measurements than has been recognized until now. We have found that the peaks of the spectra in the Stoner region are the image of a few interband densities of Stoner states of the material. These are very sensitive to the electronic structure of the material and illustrate very clearly the allowed and forbidden interband excitations. The important band couples in SPEELS are determined by the average weight of the squared matrix elements of the pendant Stoner excitations. In this respect, umklapp processes play a most fundamental role. Our model also predicts that optical spin waves should be excited in SPEELS experiments, with umklapp scattering providing the necessary oscillator strength. Our results allow us also to explain several of the features observed in SPEELS spectra. From the theoretical point of view, ab initio calculations of the interband densities of Stoner states and matrix elements $`W_{nn^{}}`$ would provide closer look to the details of the mechanisms behind SPEELS. The differences between different ferromagnets, like Fe and Ni, could also be better understood. From the experimental point of view, measurements with higher resolution would be desirable, particularly for the detection of spin waves. We think our results provide a good starting point for those further studies.
###### Acknowledgements.
R.S. would like to thank the members of the Department of Applied Physics at Chalmers University of Technology for their hospitality during a stay in which part of this work was done. This project was supported by the Swedish Natural Science Research Council. R.S. would also like to thank Suk Joo Youn and Ove Jepsen for LMTO-ASA data for Fe, and Anthony Paxton for providing him with useful literature. Many thanks in particular to Prof. A. J. Freeman for encouraging discussions.
|
warning/0006/hep-th0006130.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
####
The physical and mathematical aspects of the vector equations $`\stackrel{}{}\times \stackrel{}{V}=k\stackrel{}{V}`$($`\stackrel{}{V}`$ is a vector field and $`k`$ is a positive constant) have frequently been analyzed. Particularly, plasma and astrophysical plasma physics has pointed out the existence of stationary eletromagnetic waves that seems to be the most important aspect of this equation. Particular solutions have been found for the classical electrodynamics vector potential with the form \[1-8\]
$$\stackrel{}{A}=a[i\mathrm{sin}kz+j\mathrm{cos}kz]cos\omega t$$
(1.1)
The associated electric and magnetic fields are:
$`\stackrel{}{E}`$ $`=`$ $`{\displaystyle \frac{1}{c}}{\displaystyle \frac{\stackrel{}{A}}{t}}=ka[i\mathrm{sin}kz+\widehat{j}\mathrm{cos}kz]\mathrm{sin}\omega t`$ (1.2)
$`\stackrel{}{B}`$ $`=`$ $`\stackrel{}{}\times \stackrel{}{A}=ka[i\mathrm{sin}kz+j\mathrm{cos}kz]\mathrm{cos}\omega t`$ (1.3)
The fields (2) and (3) satisfy Maxwell equations, as well as, the usual vacuum free-wave equation. Moreover, fields $`\stackrel{}{E}`$, $`\stackrel{}{B}`$ and $`\stackrel{}{A}`$ are parallel everywhere. It is interesting to point out, these electromagnetic waves do not propagate energy, having a null Poynting vector. From the $`\stackrel{}{}\times \stackrel{}{A}=k\stackrel{}{A}`$, we get the equation:
$$^2\stackrel{}{A}+k^2\stackrel{}{A}=0$$
(1.4)
with equation (1) as a possible solution.
We argue that if astrophysical plasma is considered, there must be some restrictions to get the parallel $`\stackrel{}{A}`$, $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ fields solutions reported by Brownstein et. al. .
This work shows that it is not always possible to get these solutions if a strong gravitational background is considered. Gravitation breaks the fields parallelism. So the associated eletromagnetic wave does not have a null Poynting vector anymore and propagates energy in a gravitational background. Then, looking for equation like (1.4) for gravitational field and a particular solution, we analyze of a possible stationary gravitational wave. It could be new gravitational waves, different from the general relativity gravitational waves (taken from linearization of Einstein equations). Experimentally, this stationary gravitational waves may be found in black-hole distributions.
Finally, using electromagnetic and gravitational gauge theory symmetries we analyse stationary quantum gravitational wave. Here we consider gravitation as a $`4`$-dimensional effective theory.
## 2 Parallel $`\stackrel{}{A}`$, $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ fields in Plasma and Astrophysical Plasma
####
The fields $`\stackrel{}{A}`$, $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ are parallel in vaccum or, under some conditions, in plamas, what should follow if a gravitational background is taken into account? Do Brownstein fields remain parallel? and does the associated electromagnetic wave propagate energy?
In order to check these questions we consider a gravitational and electromagnetic coupling by the action:
$$๐ฎ=\sqrt{g}\left(\frac{1}{4}F_{\mu \nu }F^{\mu \nu }\right)d^4x$$
(2.5)
In this gravitational background the Maxwell inhomogeneous equations are:
$$๐_\mu F^{\mu \nu }=๐ฅ^\nu $$
(2.6)
where
$$๐_\mu F^{\mu \nu }=_\mu F^{\mu \nu }+\mathrm{\Gamma }_{\mu \lambda }^\mu F^{\lambda \nu }+\mathrm{\Gamma }_{\mu \lambda }^\nu F^{\mu \lambda }=๐ฅ^\nu $$
(2.7)
and Maxwell homogeneous equations are:
$$๐_\mu F_{\nu \rho }+๐_\nu F_{\rho \mu }+๐_\rho F_{\mu \nu }=0$$
(2.8)
The connections terms cancel each other such that this last equation is the usual Maxwell homogeneous equation:
$$_\mu F_{\nu \rho }+_\nu F_{\rho \mu }+_\rho F_{\mu \nu }=0.$$
(2.9)
To solve inhomogeneous equations (8) we adopt the F.R.W. cosmological metric:
$$dS^2=dt^2a_{(t)}^2\left\{\left(1Ar^2\right)^1dr^2+r^2d\theta ^2+r^2\mathrm{sin}^2\theta d\phi ^2\right\}$$
(2.10)
At the same time, the second term of the covariant derivative equation (7) may be written appropriately, with $`\mathrm{\Gamma }_{\mu \lambda }^\mu `$ given as
$$\mathrm{\Gamma }_{\mu \lambda }^\mu =\frac{}{x^\lambda }\mathrm{log}\sqrt{\stackrel{~}{g}},$$
(2.11)
where $`\stackrel{~}{g}`$ is the metric determinant.
Then Maxwell equations (7) and (8) are explicitly given by:
$$\stackrel{}{}\stackrel{}{E}g\stackrel{}{}f\stackrel{}{E}=\rho _{(\stackrel{}{x})},$$
(2.12)
$$\stackrel{}{}\stackrel{}{B}=0,$$
$$\stackrel{}{}\times \stackrel{}{E}=\frac{\stackrel{}{B}}{t},$$
$$\stackrel{}{}\times \stackrel{}{B}=\stackrel{}{J}(\stackrel{}{x})+\frac{\stackrel{}{E}}{t}g\frac{f}{t}\stackrel{}{E}+g\stackrel{}{}f\times \stackrel{}{B}\mathrm{\Gamma }_{\mu \lambda }^iF^{\mu \lambda }.$$
Without electromagnetic sources, that is, for $`\rho =0`$ and $`\stackrel{}{J}=0`$ we get โfreeโ electromagnetic fields equations in a graviational background:
$$\stackrel{}{}\stackrel{}{E}=g\stackrel{}{}f\stackrel{}{E},$$
(2.13)
$$\stackrel{}{}\stackrel{}{B}=0,$$
(2.14)
$$\stackrel{}{}\times \stackrel{}{E}=\frac{\stackrel{}{B}}{t},$$
(2.15)
$$\stackrel{}{}\times \stackrel{}{B}=\frac{\stackrel{}{E}}{t}g\frac{f}{t}\stackrel{}{E}+g\stackrel{}{}f\times \stackrel{}{B}\mathrm{\Gamma }_{\mu \lambda }^iF^{\mu \lambda }$$
(2.16)
where. $`g={\displaystyle \frac{\sqrt{1Ar^2}}{a^3r^2\mathrm{sin}\theta }}`$ and $`f={\displaystyle \frac{a^2r^2\mathrm{sin}\theta }{\sqrt{1Ar^2}}}`$.
Functions $`g`$ and $`f`$ may have $`A=+1,0,1`$ each value represents the associated curvature of F.R.W. spatial metric section.
The electric field wave equation in a gravitational background is obtained from eq. (15) as
$`^2\stackrel{}{E}`$ $``$ $`{\displaystyle \frac{^2\stackrel{}{E}}{t^2}}=\stackrel{}{}\left(g\stackrel{}{}f\stackrel{}{E}\right){\displaystyle \frac{g}{t}}{\displaystyle \frac{f}{t}}\stackrel{}{E}g{\displaystyle \frac{^2f}{t^2}}\stackrel{}{E}g{\displaystyle \frac{f}{t}}{\displaystyle \frac{\stackrel{}{E}}{t}}+`$
$`+`$ $`{\displaystyle \frac{g}{t}}\stackrel{}{}f\times \stackrel{}{B}+g{\displaystyle \frac{}{t}}\stackrel{}{}f\times \stackrel{}{B}+g\stackrel{}{}f\times {\displaystyle \frac{\stackrel{}{B}}{t}}{\displaystyle \frac{}{t}}\left(\mathrm{\Gamma }_{\mu \beta }^iF^{\mu \beta }\right).`$
The magnetic field wave equation in a gravitational background is obtained from eq. (16) as
$$^2\stackrel{}{B}\frac{^2\stackrel{}{B}}{t^2}=\stackrel{}{}\times \left(g\frac{f}{t}\stackrel{}{E}g\stackrel{}{}f\times \stackrel{}{B}+\mathrm{\Gamma }_{\mu \beta }^iF^{\mu \beta }\right).$$
(2.18)
The last term, $`\mathrm{\Gamma }_{\mu \lambda }^\nu F^{\mu \lambda }`$, is identically zero. But we still retain it for esthetic completeness of equation (2.7). At the same time, equations (2.17) and (2.18) reproduce the correponding vacuum equations if gravitation is ignored.
Using equation (2.17) and (2.18) it can be shown that there is at least, one solution in which fields $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ are not parallel anymore. For example, putting field $`\stackrel{}{B}`$ as given by (2.3), in equation (2.18), and writting the gradient $`\stackrel{}{}f`$ as:
$$\stackrel{}{}f=\widehat{r}_0\frac{f}{r}+\widehat{\theta }_0\frac{f}{\theta }+\widehat{\phi }_0\frac{f}{\phi },$$
(2.19)
it is easy to verify that equation (2.18) becomes
$$\frac{f}{t}\stackrel{}{E}=\stackrel{}{}f\times \stackrel{}{B}\stackrel{}{E}=\frac{\stackrel{}{}f\times \stackrel{}{B}}{{\displaystyle \frac{f}{t}}}$$
(2.20)
Calculating $`{\displaystyle \frac{f}{t}}`$, $`\stackrel{}{}f`$, $`\stackrel{}{}f\times \stackrel{}{B}`$ and taking explicitly the unity vectors $`\widehat{r}_0`$ and $`\widehat{\theta }_0`$
$`\widehat{r}_0`$ $`=`$ $`\widehat{i}\mathrm{sin}\theta \mathrm{cos}\phi +\widehat{j}\mathrm{sin}\theta \mathrm{sin}\phi +\widehat{k}\mathrm{cos}\theta `$ (2.21)
$`\widehat{\theta }_0`$ $`=`$ $`\widehat{i}\mathrm{cos}\theta \mathrm{cos}\phi +\widehat{j}\mathrm{cos}\theta \mathrm{sin}\phi \widehat{k}\mathrm{sin}\theta `$
we can find,
$$\widehat{B}=ka\left[i\mathrm{sin}(kz)+\widehat{j}(kz)\right]\mathrm{cos}(\omega t),$$
(2.22)
$`\stackrel{}{E}`$ $`=`$ $`\widehat{i}\left(\mathrm{sin}\theta G_{(r,t,\theta )}\mathrm{cos}\theta F_{(r,t)}\right)ka\mathrm{cos}(kz)\mathrm{cos}(\omega t)+`$
$`+`$ $`\widehat{j}\left(\mathrm{cos}\theta F_{(r,t)}\mathrm{sin}\theta G_{(r,t,\theta )}\right)ka\mathrm{sin}(kz)\mathrm{cos}(\omega t)+`$
$`+`$ $`\widehat{k}[(\mathrm{sin}\theta \mathrm{cos}\phi F_{(r,t)}+\mathrm{cos}\theta \mathrm{cos}\phi G_{(r,t,\theta )})ka\mathrm{cos}(kz)\mathrm{cos}(\omega t)+`$
$``$ $`\left(\mathrm{sin}\theta \mathrm{sin}\phi F_{(r,t)}\mathrm{cos}\theta \mathrm{sin}\phi G_{(r,t,\varphi )}\right)ka\mathrm{sin}(kz)\mathrm{cos}(\omega t)]`$
where functions $`F_{(r,t)}`$ and $`G_{r,t,\theta )}`$ are given by:
$`F_{(r,t)}`$ $`=`$ $`{\displaystyle \frac{2a}{3\dot{a}r}}+{\displaystyle \frac{Aar}{3\dot{a}(1Ar^2)}}\text{and}`$
$`G_{(r,t,\theta )}`$ $`=`$ $`{\displaystyle \frac{a\text{cotg}\theta }{3\dot{a}r}}.`$
The new electrital field $`\stackrel{}{E}`$ given by (2.23) is completely different from the Brownstein electrical field (1.2). Now, fields $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ are perpendicular vectors, as we can see from equation (2.20). Gravitational backgound breaks $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ parallelism and so the Poynting vector $`\stackrel{}{S}=\stackrel{}{E}\times \stackrel{}{B}`$ is not zero anymore.
This way, the associated electromagnetic wave propagates energy.
It is possible the electric field as the Brownstein electric field given in eq. (1.2) is kept and a new particular solution for magnetic $`\stackrel{}{B}`$ field in a gravitational background is obtained. We believe, this new magnetic field differs from the Brownstein magnetic field given in eq. (1.3) and the associated electromagnetic wave propagates energy.
It is not always possible to have $`\stackrel{}{E}`$, $`\stackrel{}{B}`$ and $`\stackrel{}{A}`$ as parallel fields if a gravitational background is taken into account. On the other hand, even in a gravitational background, it might be possible to claim parallel $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ fields, writing $`\stackrel{}{E}=\chi \stackrel{}{B}`$ in equations (2.17) and (2.18) and to get conditions that the function $`\chi `$ must satisfy.
## 3 Stationary Gravitational Waves
####
Now, considering the similarities of electromagnetic, gravitational and linearized Einstein gravitation theories, we analyze a possibility of stationary gravitational wave.
According to the phenomenological view, black-hole distributions may be nodes of gravitational waves propagating in space. It would be sufficient for two black-holes separated by large distances, if the space-time curvature of the strong gravitational fields of the black-holes is neglected. So the space-time of one black-hole is assymptotically flat to the other. Then each black-holes is like a โstring nodeโ and this mechanism may confine gravitational waves.
The following equation may reproduce these waves in nature:
$$R_{\mu \nu }=\kappa \mathrm{\Lambda }h_{\mu \nu }$$
(3.1)
where
$$g_{\mu \nu }=\eta _{\mu \nu }+\kappa h_{\mu \nu }.$$
(3.2)
Here $`\kappa `$ is the Newton gravitational constant, $`\mathrm{\Lambda }`$ is the cosmological constant, $`h_{\mu \nu }`$ is the stationary gravitational perturbation in a Minkowski โBackgroundโ. Then the stationary gravitational pertubation is given by
$$_\beta _\nu h_\mu ^\beta +_\beta _\mu h_\nu ^\beta \mathrm{}h_{\mu \nu }_\mu _\nu h_\beta ^\beta =\mathrm{\Lambda }h_{\mu \nu }.$$
(3.3)
This equation is formally similar to the electromagnetic case, eq. (1.4).
This equation may have a solution like
$$h_{\mu \nu }=C_{\mu \nu }(z)f(t),$$
(3.4)
where
$$h_{\mu \nu }=\left[\begin{array}{cccc}A_{00}& 0& 0& 0\\ 0& A_{11}& A_{12}& 0\\ 0& A_{12}& A_{11}& 0\\ 0& 0& 0& A_{00}\end{array}\right]e^{i\stackrel{~}{k}z}\mathrm{cos}\omega t,$$
(3.5)
where $`A_{00}`$, $`A_{11}`$ and $`A_{12}`$ are free constants and $`\stackrel{~}{k}=\sqrt{\mathrm{\Lambda }\omega ^2}`$ is the wave-vector for the case $`\mathrm{\Lambda }\omega ^2>0`$.
The dispersion relation suggests a long wave-length, $`\lambda `$, for stationary gravitational waves in nature, since its frequency is small.
The lagrangean for the pertuberd field is
$`_h^{(\mathrm{\Lambda })}`$ $`=`$ $`{\displaystyle \frac{1}{4}}_\rho h_{\mu \nu }^\rho h^{\mu \nu }+{\displaystyle \frac{1}{4}}_\rho h^\rho h{\displaystyle \frac{1}{2}}_\rho h^{\rho \mu }_\mu h+{\displaystyle \frac{1}{2}}^\rho h_{\rho \mu }_\nu h^{\mu \nu }+`$ (3.6)
$``$ $`{\displaystyle \frac{1}{4}}\mathrm{\Lambda }h^{\mu \nu }h_{\mu \nu }+{\displaystyle \frac{1}{8}}\mathrm{\Lambda }h^2.`$
ยฟFrom this lagrangean we get the energy-momentum tensor $`T_{\mu \nu }`$ and, using the equation of motion with $`_\mu T^{\mu \nu }=0`$. The stress tensor is not symmetric. Then $`T_{\mu \nu }`$ is
$`T_{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{1}{4}}_\mu _\nu h{\displaystyle \frac{1}{2}}_\mu h_{\alpha \beta }_\nu h^{\alpha \beta }{\displaystyle \frac{1}{2}}_\nu h_{\mu \beta }^\beta h{\displaystyle \frac{1}{2}}^\beta h_{\mu \beta }_\nu h+`$ (3.7)
$`+`$ $`^\beta h_{\beta \sigma }_\nu h_\mu ^\sigma \eta _{\mu \nu }_h^{(\mathrm{\Lambda })}.`$
A system of black-hole distribution may confine stationary gravitational waves.
## 4 Gravitation as a Gauge Theory-Gravitons
####
Now, we intend to attack the last problem according to quantum view. Gravitation is considered a gauge theory.
Classically, stationary gravitational waves between two points $`x`$ and $`y`$ have been obtained. According to the quantum view, this problem may have graviton exchanges. The situation is similar to that of point charges interacting by photon exchanges.
This way, the quantum version of the problem of stationary gravitational waves between two points $`x`$ and $`y`$ is now described by graviton creation and annihilation at these same points.
For convecinece, we parametrize the field $`h_{\mu \nu }`$ as
$$H_\nu ^\alpha =h_\nu ^\alpha \frac{1}{2}\delta _\nu ^\alpha h.$$
It is simple to verify that $`H_\alpha ^\alpha =H=h`$ where $`h`$ is the trace of $`h_{\mu \nu }`$. With this new $`h_{\mu \nu }`$, the Lagrangian (29) may be written as
$`_H^{(\mathrm{\Lambda })}`$ $`=`$ $`{\displaystyle \frac{1}{2}}_\rho H_{\mu \nu }^\rho H^{\mu \nu }+{\displaystyle \frac{1}{4}}_\rho H^\rho H{\displaystyle \frac{1}{2}}^\rho H_{\rho \nu }_\mu H^{\mu \nu }+{\displaystyle \frac{1}{2}}^\rho H_{\rho \mu }_\nu H^{\mu \nu }+`$ (4.8)
$``$ $`{\displaystyle \frac{1}{2}}\mathrm{\Lambda }H^{\mu \nu }H_{\mu \nu }+{\displaystyle \frac{1}{4}}\mathrm{\Lambda }H^2.`$
Finally, we integrate the lagrangean (31) by parts and get:
$`_H^{(\mathrm{\Lambda })}`$ $`=`$ $`{\displaystyle \frac{1}{2}}H^{\mu \nu }\mathrm{}H^{\mu \nu }{\displaystyle \frac{1}{4}}H\mathrm{}H{\displaystyle \frac{1}{2}}H^{\mu \nu }_\mu _\alpha H_\nu ^\alpha {\displaystyle \frac{1}{2}}H^{\mu \nu }_\nu _\alpha H_\nu ^\alpha +`$ (4.9)
$``$ $`{\displaystyle \frac{1}{2}}\mathrm{\Lambda }H^{\mu \nu }H_{\mu \nu }+{\displaystyle \frac{1}{4}}\mathrm{\Lambda }H^2.`$
Then, it is possible to write this expression in a bilinear form:
$$_H^{(\mathrm{\Lambda })}=\frac{1}{2}H^{\mu \nu }\mathrm{\Theta }_{\mu \nu ,\kappa ,\lambda }H^{\kappa ,\lambda },$$
(4.10)
where operator $`\mathrm{\Theta }_{\mu \nu ,\kappa \lambda }`$ has the following form in terms of Barnes-Rivers spin projection operators:
$`\mathrm{\Theta }_{\mu \nu ,\kappa \lambda }`$ $`=`$ $`(\mathrm{}\mathrm{\Lambda })P^{(2)}\mathrm{\Lambda }P_m^{(1)}+{\displaystyle \frac{5}{2}}(\mathrm{}\mathrm{\Lambda })P_s^{(0)}{\displaystyle \frac{(\mathrm{\Lambda }+3\mathrm{})}{2}}P_\omega ^{(0)}`$
$`+`$ $`{\displaystyle \frac{\sqrt{3}}{2}}(\mathrm{\Lambda }\mathrm{})P_\omega ^{(0)}+{\displaystyle \frac{\sqrt{3}}{2}}(\mathrm{\Lambda }\mathrm{})P_{\omega s}^{(0)}.`$
The inverse operator $`\mathrm{\Theta }_{\mu \nu ,\kappa \lambda }^1`$ has the following form:
$$\mathrm{\Theta }_{\mu \nu ,\kappa \lambda }^1=\left[XP^{(2)}+YP_m^{(1)}+ZP_s^{(0)}+WP_\omega ^{(0)}+RP_{sw}^{(0)}+SP_{\omega s}^{(0)}\right]_{\mu \nu ,\kappa \lambda }.$$
(4.12)
To get the coefficients $`X,Y\mathrm{}S`$ it is sufficient to use Barnes-Rivers projection operators and its multiplicative table, such that:
$$\mathrm{\Theta }^{\rho \sigma }_{\mu \nu }\mathrm{\Theta }_{\rho \sigma ,\kappa \lambda }^1=(I)_{\mu \nu ,\kappa \lambda }=\left(P^{(2)}+P_m^{(1)}+P_s^{(0)}+P_\omega ^{(0)}\right)_{\mu \nu ,\kappa \lambda }.$$
(4.13)
For $`D=4`$ these coefficients have the following form
$$X=\frac{1}{\mathrm{\Lambda }\mathrm{}},Y=\frac{1}{\mathrm{\Lambda }},Z=\frac{\mathrm{\Lambda }+3\mathrm{}}{\mathrm{\Lambda }^2+8\mathrm{\Lambda }\mathrm{}9\mathrm{}^2},$$
(4.14)
$$W=\frac{5}{\mathrm{\Lambda }9\mathrm{}},R=\frac{\sqrt{3}}{\mathrm{\Lambda }+\mathrm{}},S=\frac{\sqrt{3}}{\mathrm{\Lambda }+9\mathrm{}}.$$
We get the corresponding propagator by the following functional generator:
$$W[T_{\rho \sigma }]=\frac{1}{2}d^4xd^4yT^{\mu \nu }\mathrm{\Theta }_{\mu \nu ,\kappa \lambda }^1T^{\kappa \lambda }.$$
The propagator is written explicitly as:
$$T[h_{\mu \nu (x)};h_{\kappa \lambda (x)}]=i\mathrm{\Theta }_{\mu \nu ,\kappa \lambda }^1\delta ^4(xy).$$
(4.15)
## 5 Unitarity of the Model
####
Now we discuss the tree-level unitarity of the model. Coupling the propagator and external current, $`T^{\mu \nu }`$ we analyze the poles of the current amplitude and the imaginary part of its residue. The current amplitude is given by:
$$๐=T^{\mu \nu }(\stackrel{}{k})T[h_{\mu \nu }(\stackrel{}{k});h_{\kappa \lambda }(\stackrel{}{k})]T^{\kappa \lambda }(\stackrel{}{k}).$$
(5.16)
Using equations (4.12), (4.14) and (4.15) we get the current amplitude. And considering the imaginary part of the residue at the poles, it is simple to verify that, due to the transverse condition, only operators $`P^{(2)}`$ and $`P_s^{(0)}`$ remain. And that it is a symmetry, not necessarily a gauge symmetry:
$$\omega _{\mu \nu }T^{\mu \nu }=0.$$
(5.17)
Analyzing the poles of the sector with spin-2 and the sector with spin-$`0`$ in the momentum space we get:
$$X=\frac{1}{k^2+\mathrm{\Lambda }}$$
(5.18)
$$Z=\frac{3k^2\mathrm{\Lambda }}{\mathrm{\Lambda }^28\mathrm{\Lambda }k^29(k^2)^2}.$$
This way we have for the sector with spin-$`2`$, a tachyon pole:
$$k^2=\mathrm{\Lambda }$$
(5.19)
and for the sector with spin-$`0`$, we have two poles, one being a tachyon pole
$`k^2`$ $`=`$ $`{\displaystyle \frac{1}{9}}\mathrm{\Lambda },`$ (5.20)
$`k^2`$ $`=`$ $`\mathrm{\Lambda }.`$
Using (5.16) and considering the $`k^2`$ pole analysis we can verify the tree-level unitarity of the model. But we want spin-$`2`$ gravitons, so we redefine $`\mathrm{\Lambda }`$ as $`(\mathrm{\Lambda })`$ and, so, for the spin-$`2`$ sector, we have
$$k^2=\mathrm{\Lambda }.$$
(5.21)
For the spin-$`0`$ sector, we have, respectively:
$$k^2=\frac{1}{9}\mathrm{\Lambda },$$
(5.22)
$$k^2=\mathrm{\Lambda }.$$
Now we have one non tachyonic pole for the spin-$`2`$ sector, one non tachyonic and one tachyonic pole for the spin-$`0`$ sector.
And we find
$$ImRes๐|_{k^2=\mathrm{\Lambda }}>0$$
(5.23)
Thus the massive excitation is, in fact, a dynamical degree of freedom. The theory does not have negative norm states. And this is the unitarity requisite to have the required assymptotic behavior. From (41), it is simple to verify that the propagator is proportional to $`{\displaystyle \frac{1}{k^4}}`$, and therefore the $`4D`$ model is not renormalizable.
So we conclude that the model discussed is causal, has tree-level unitarity and is not renormalizable by power counting. The proposed model has to be understood as an โeffective gravitation theoryโ with a physical massive degree of freedom $`(k^2=\mathrm{\Lambda })`$.
### Acknowledgements:
####
The authors want to thank G.O. Pires for his comments and technical discussions and S.A. Diniz for his work on typing. We also thank Prof. Berth Schrรถer for many discussions during his visit to the Department of Physics of UFES. We also would like to thank the Department of Physics, University of Alberta for their hospitality. Finally we thank Conselho Nacional de Pesquisa โ CNPq โ for the financial support.
|
warning/0006/hep-ph0006090.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Recent studies of precision measurements of neutrino mixing parameters and neutrino masses focused on high intensity beams at neutrino factories, these offer unique advantages compared to conventional wide band neutrino beams. The development of a neutrino factory is however very cost intensive and still in its initial stage. For this reason it was recently emphasized that further studies of conventional wide band neutrino beams should be pursued . One of the major advantages of a neutrino factory oscillation experiment would be the possibility to measure the appearance oscillation channel $`\overline{\nu }_e\overline{\nu }_\mu `$ (wrong sign muons) in very long baseline experiments. Using appearance rates enables the determination of the small mixing angle $`\theta _{13}`$, $`|\mathrm{\Delta }m_{31}^2|`$ as well as the sign of $`\mathrm{\Delta }m_{31}^2`$, which discriminates between the two possible mass ordering schemes . An analysis of the appearance rates requires that the detector can discriminate the charges of $`\mu ^+`$ and $`\mu ^{}`$ at a sufficient level in order to separate wrong sign muons from the large amount of right sign muons produced in the disappearance channel $`\nu _\mu \nu _\mu `$. Therefore, large magnetized iron detectors usually are the detection system of choice. In it was however shown that the ability to determine $`\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$ does not necessarily depend on the capability of charge identification. The information on $`\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$ is also contained in the disappearance channel $`\nu _\mu \nu _\mu `$ and can thus be extracted from the $`\nu _\mu `$ rates without charge identification. If $`\theta _{13}`$ is not too small a measurement is possible in this channel with a precision which is comparable to the appearance channel. Furthermore, such a measurement would even be possible with conventional wide band neutrino beams, if a sufficiently high neutrino event rate is achieved, which is big enough to limit the statistical error, and if the systematic errors on the beam flux are under control. We show in this paper that conventional neutrino beams (consisting only of $`\nu _\mu `$) in combination with large water or ice Cherenkov detectors (Neutrino Telescopes without charge identification) can be used and give remarkable results. In particular we use as a prototype scenario a CNGS-type beam and an AMANDA-like detector , for which the neutrino event rates are comparable to those of proposed neutrino factory experiments. We discuss the problems which arise in measurements of neutrinos in the energy threshold region of Neutrino Telescopes and suggest using beam pulse timing information and neutrino direction information to reduce the background from atmospheric muons. Finally, we perform a numerical analysis of the physics potential of this type of experiment and show that precision measurements of the leadings oscillation parameters as well as the determination of $`\mathrm{sin}^22\theta _{13}`$, the test of the MSW-effect and the determination of the sign of $`\mathrm{\Delta }m^2`$ are possible.
## 2 Three Neutrino Oscillations in Matter
The basic mechanism which allows the extraction of the sign of $`\mathrm{\Delta }m_{31}^2`$ comes from coherent forward scattering of electron neutrinos in matter (MSW-effect) which leads to effective masses and mixings different from vacuum. In the approximation where the solar $`\mathrm{\Delta }m^2`$ is ignored compared to the atmospheric $`\mathrm{\Delta }m^2`$, i.e. $`\mathrm{\Delta }m_{21}^2=0`$ they are given by $`\mathrm{\Delta }m_{31,m}^2=\mathrm{\Delta }m^2C_\pm `$, $`\mathrm{\Delta }m_{32,m}^2=\mathrm{\Delta }m^2[(C_\pm +1)+A]/2`$, $`\mathrm{\Delta }m_{21,m}^2=\mathrm{\Delta }m^2[(C_\pm 1)A]/2`$ and $`\mathrm{sin}^22\theta _{13,m}=\mathrm{sin}^22\theta _{13}/C_\pm ^2`$ where $`C_\pm =[(A/\mathrm{\Delta }m^2\mathrm{cos}2\theta )^2+\mathrm{sin}^22\theta ]^{1/2}`$ and $`A=2EV=\pm 2\sqrt{2}G_FY\rho E/m_n`$. $`\theta _{23}`$ is not changed in the approximation $`\mathrm{\Delta }m_{21}^2=0`$ (we follow in this work the notation of ). The size of the matter corrections for a given neutrino species depends on the sign of $`\mathrm{\Delta }m_{13}^2`$ which enters into $`C_\pm `$. The resulting modification of the muon disappearance probability in matter therefore depends on the sign of $`\mathrm{\Delta }m_{13}^2`$:
$`P(\nu _\mu \nu _\mu )=1`$ $``$ $`\mathrm{sin}^22\theta _{23}\mathrm{sin}^2\theta _{13,m}\mathrm{sin}^2(\mathrm{\Delta }_{31,m})`$
$``$ $`\mathrm{sin}^4\theta _{23}\mathrm{sin}^22\theta _{13,m}\mathrm{sin}^2(\mathrm{\Delta }_{31,m})`$
$``$ $`\mathrm{sin}^22\theta _{23}\mathrm{cos}^2\theta _{13,m}\mathrm{sin}^2(\mathrm{\Delta }_{31,m}).`$
For antineutrinos, the matter induced modifications of the disappearance probability correspond to opposite sign of $`\mathrm{\Delta }m_{31}^2`$ for neutrinos. This can be used to amplify the signature of matter effects if $`\nu _\mu `$ and $`\overline{\nu }_\mu `$ beams can be compared. The biggest effect from the sign of $`\mathrm{\Delta }m_{13}^2`$ will be seen at the MSW resonance energy (in the Earth $``$ 10-15 GeV) and the sensitivity to matter effects will thus be best when the maximum of the neutrino spectrum is suitably adjusted and the detector is sensitive in this energy region. For a more detailed treatment see .
The differential muon event rates in the detector are:
$$\frac{dn_\mu }{dE}=\underset{normalization}{\underset{}{10^9N_AN_{\text{kT}}ฯต_\mu }}\underset{flux}{\underset{}{\left(\frac{L_{\mathrm{near}}}{L_{\mathrm{far}}}\right)^2f_{\mathrm{near}}(E)\sigma (E)}}\underset{oscillation}{\underset{}{P_{\nu _\mu \nu _\mu }(E)}}.$$
(2)
Here $`10^9N_A`$ is the number of nucleons per kiloton in the detector, $`N_{\text{kT}}`$ is the detector size in kilotons, $`ฯต_\mu `$ is the detection efficiency and $`\sigma `$ is the charged current neutrino cross section per nucleon. The neutrino beam flux $`f_{\mathrm{near}}`$, which we assume to be monitored at the near detector, must be corrected by the geometrical suppression factor $`(L_{\mathrm{near}}/L_{\mathrm{far}})^2`$, where $`L_{\mathrm{near}}`$ and $`L_{\mathrm{far}}`$ are the distances to the near and far detectors. Total rates are then obtained by integrating these differential rates from the threshold to the maximum neutrino energy.
## 3 Experimental Issues
Next we turn to the physics potential of large water or ice Cherenkov detectors as components of very long baseline neutrino oscillation experiments. Several experiments of this type which consist of photomultipliers which are attached to vertical strings are under construction or already taking data (AMANDA / ICECUBE , ANTARES , BAIKAL and NESTOR ). For the proposed ICECUBE array, the vertical spacing is about 15 m, and the 80 strings of 1 km length are 100 m away from each other. These Cherenkov detectors allow detector masses up to $`10^5`$ kt which is a factor 100 more than the most ambitious ideas proposed for a megaton Super-Kamiokande-like water Cherenkov detector. These detectors, which conventionally are called โNeutrino Telescopesโ, are primarily built in order to search for very high energy cosmic and atmospheric neutrinos. In combination with neutrino beams, some comments are in order concerning measurement of neutrinos very near the threshold energy of the detector (20 to 100 GeV). For very long baseline oscillation experiments it is important to achieve a detector energy threshold of 10 GeV to 20 GeV and reasonable energy resolution of about 10 GeV. This low neutrino energy is needed to take advantage of the MSW-effects which enhances sub-leading effects from $`\theta _{13}`$ and from the sign of $`\mathrm{\Delta }m_{31}^2`$. The usually quoted energy thresholds for the detectors mentioned above are a result of the geometry of the photo multiplier (PM) array, the optical properties of ice or sea water and the specific energy loss of muons per traveled distance ($`0.2\mathrm{GeV}/\mathrm{m}`$). Only upward moving muons are unambiguously assigned to neutrino interactions and the ability to reconstruct the particle track with sufficient precision is crucial in order to reject the huge background from downward going muons. Typically, hits in 7 or more PMs are requested for proper track reconstruction. For AMANDA, the most simple majority trigger requests 8 to 16 hit PMs, and this sets the energy threshold to 30-50 GeV. By triggering on certain hit patterns like hits in 5 neighboring (or nearly neighboring) PMs with some of them having high amplitude, one can select muons close to a string, with lower energy threshold (and of course paying with a reduced effective volume). Still, the background of downward muons dwarfs the signal, since the tracking error obtained from 5 hits is rather large. The situation is however different, when a neutrino beam is used as a source: The background can be reduced by five to six orders of magnitude, if beam pulse timing information<sup>1</sup><sup>1</sup>1The duty cycle of wide band neutrino beams is about $`1:10^6`$, that of neutrino factories may reach $`1:10^3`$ . The absolute timing accuracy of events is about 100 ns (GPS accuracy). and beam direction information<sup>2</sup><sup>2</sup>2In the case discussed here, the neutrino beam will meet the detector under a small angle to the direction of the strings. is used. Independently a lower bound on the threshold is imposed by the requirement to distinguish hadronic showers which are produced by neutral current reactions of all neutrino flavors. The length of such a shower grows as $`\mathrm{log}E`$; the track length of a muon grows as $`E`$; therefore a cut on the minimum track-length is used which corresponds to an energy threshold of 10 GeV to 20 GeV, depending on the optical properties of the detector medium, and yields an energy resolution in the same range.
These arguments suggest strongly that Neutrino Telescopes can be used as large mass detectors for very long baseline neutrino experiments, but this idea requires certainly further investigation. For our work we assume a prototype detector with an effective mass of 5000 kt. This seems to be a reasonable estimate for the effective volume of AMANDA and ANTARES with respect to muons in the range of 10-20 GeV. The corresponding ICECUBE volume might come close to 10 to 100 Mt. The assumed energy threshold is 15 GeV and the energy resolution is 10 GeV . To illustrate the enormous potential of such a detector, we use as source a conventional wide band neutrino beam like NuMI or CNGS. The CNGS beam consists mainly of $`\nu _\mu `$ with an admixture of about 2% $`\overline{\nu }_\mu `$, 0.8% $`\nu _e`$ and 0.05% $`\overline{\nu }_e`$ and the number of additional muons expected from oscillated $`\nu _e`$โs in the beam is small and negligible. The mean energy of the beam lies in the range of 20 GeV to 30 GeV, with a long tail towards higher energies. In principle it is also possible to get a $`\overline{\nu }_\mu `$-beam by reversing the current in the lens system; this would result however in an smaller flux of about 75% and larger admixtures of other neutrinos.
## 4 Results
For our analysis we assume a 5000 kt detector with energy threshold 15 GeV, energy resolution 10 GeV (corresponding to four bins) and $`ฯต_\mu =100\%`$ detection efficiency. As neutrino source we used the CNGS flux spectrum for $`4.510^{19}`$ protons on target . We numerically calculate the charged current rate spectrum at the detector according to eq. 2. To compute the transition probability we integrate eq. 2 over the full Stacey matter density profile of the Earth . We then impose Poisson fluctuations on the rates and re-extract the oscillation parameters with a maximum likelihood method (for details see ). Systematic errors of the beam flux and backgrounds are assumed to be small and are not included. The calculations are performed in the approximation $`\mathrm{\Delta }m_{21}^2=0`$, i.e. the solar mass squared splitting is ignored which is here a very good approximation. We investigate two different baselines. First 6500 km (approximately FermilabโANTARES), to measure $`\theta _{13}`$ and then 11200 km (CERNโAMANDA), to test the MSW-effect and to determine the sign of $`\mathrm{\Delta }m_{31}^2`$.
In fig. 1 the total event rates at a baseline of 6500 km are shown as function of the atmospheric mass squared difference $`\mathrm{\Delta }m_{31}^2`$ for different values of $`\theta _{13}`$ and for the two possible signs of $`\mathrm{\Delta }m_{31}^2`$ (labeled $``$ and $``$).
Apparently, a nonzero $`\theta _{13}`$ induces a significant depletion of the total rates only in the case of a positive sign of $`\mathrm{\Delta }m_{31}^2`$. The reason for this is that the beam consists only of muon neutrinos which show in matter MSW-resonant enhancement only for positive $`\mathrm{\Delta }m_{31}^2`$. In this case, studies of the sub-leading oscillation parameter $`\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$ give a statistically more significant result than with negative sign of $`\mathrm{\Delta }m_{31}^2`$ (see discussion below). If indeed the sign of $`\mathrm{\Delta }m_{31}^2`$ is negative, better significance would be achieved with an antineutrino beam, since antineutrinos are resonant in this case. It might, however, not be easy to cope with the systematic errors in the beam fluxes which are substantially different in the charge conjugated channel.
In order to simulate the extraction of the oscillation parameters, we performed fits to simulated spectral rates. Instead of performing a global fit to all parameters we first fitted the leading oscillation parameters $`\theta _{23}`$ and $`|\mathrm{\Delta }m_{31}^2|`$, which are, to a good approximation, independent of $`\mathrm{sin}^22\theta _{13}`$. As second step, the sub-leading parameters $`\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$ are fitted with the leading parameters fixed to the previously obtained best fit values. The result of the fit of the leading parameters for a given sample pair of parameters ($`\mathrm{sin}^22\theta _{23}=0.6`$, $`\mathrm{\Delta }m_{31}^2=4.010^3\mathrm{eV}^2`$) is shown in fig. 2. The relative $`3\sigma `$-errors for the whole parameter region which is allowed by the Super-Kamiokande experiment are between 0.5% and 10%, depending on the value of $`\mathrm{\Delta }m_{31}^2`$.
With the result of the fit of the leading parameters, it is possible to proceed to the second step: the fit of the sub-leading parameters $`\mathrm{sin}^22\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$. Fig. 3 shows the parameter region in the $`\mathrm{sin}^22\theta _{13}`$$`\mathrm{\Delta }m_{31}^2`$ plane in which the obtained simulated spectral rates are not consistent with $`\theta _{13}0`$ at 90% confidence level. The shaded area is excluded by the CHOOZ experiment . As explained above, the result differs for the two possible signs of $`\mathrm{\Delta }m_{31}^2`$. For positive $`\mathrm{\Delta }m_{31}^2`$ (solid line), a sensitivity for mixing angles down to $`\mathrm{sin}^22\theta _{13}210^3`$ is achieved. For negative $`\mathrm{\Delta }m_{31}^2`$ (dashed line), the sensitivity is worse and does not reach the $`\mathrm{sin}^22\theta _{13}10^2`$ level. In both measurements, $`\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$, the use of an antineutrino beam would increase the sensitivity to a level comparable to the case of $`\mathrm{\Delta }m_{31}^2>0`$.
In fig. 4 the $`\mathrm{sin}^22\theta _{13}`$$`|\mathrm{\Delta }m_{31}^2|`$ parameter region in which
the determination of the sign of the mass squared difference is possible at $`90\%`$ C.L for a positive sign of $`\mathrm{\Delta }m_{31}^2`$ is shown. This shows also the limit of $`\mathrm{sin}^22\theta _{13}`$ where verification of the MSW-effect is possible. Since it is very important for this measurement to include neutrinos at the MSW resonance energy ($``$ 15 GeV), the value of the threshold of the neutrino detector has a substantial influence on the obtained sensitivity. The lines shown in the plot were obtained with threshold energies of 5 GeV, 10 GeV, 15 GeV and 20 GeV. A negative $`\mathrm{\Delta }m_{31}^2`$ in combination with a $`\nu _\mu `$ beam does not produce MSW-enhanced effects in the rates and thus does not allow to determine the sign. If $`\mathrm{\Delta }m_{31}^2`$ is indeed negative, the determination of the sign of $`\mathrm{\Delta }m_{31}^2`$ and the test of matter effects requires an anti neutrino beam.
## 5 Conclusions
We studied in this paper the physics potential of large water or ice Cherenkov detectors (Neutrino Telescopes) in very long baseline accelerator neutrino oscillation experiments. In particular we have shown that conventional wide band neutrino beams pointed to detectors like AMANDA/ICECUBE, ANTARES or NESTOR give neutrino event rates at the level of $`10^4`$ to $`10^6`$ events per year. This number is comparable to rates achieved in presently proposed neutrino factory experiments. Neutrino Telescopes are primarily built for the search for ultra high energetic cosmic and atmospheric neutrinos. Conventional neutrino beams provide neutrinos of a few times 10 GeV, which is roughly at the quoted threshold of Neutrino Telescopes. To cope with this problem, we suggest to use beam pulse timing information and neutrino direction information to reduce the background produced by cosmic muons in a very effective way. The ultimate threshold is then given by limitations which arise from misidentification of hadron showers which are produced by all active neutrino flavors. We estimate that an energy threshold between 10 GeV and 20 GeV and an energy resolution between 10 GeV and 20 GeV would finally be achievable. This would open the door to a wide spectrum of interesting oscillation physics.
Our numerical study was performed in the standard three neutrino scenario under the approximation $`\mathrm{\Delta }m_{21}^2=0`$ and taking into account the full Stacey Earth density model. It demonstrates that precision measurements of the leading oscillation parameters $`\mathrm{\Delta }m_{31}^2`$ and $`\mathrm{sin}^22\theta _{23}`$ are possible. Using the CNGS beam spectrum as prototype neutrino source, the relative $`3\sigma `$-errors for the whole parameter region which is allowed by the Super-Kamiokande experiment are between 0.5% and 10%, depending on the value of $`|\mathrm{\Delta }m_{31}^2|`$. We further have demonstrated that, even though a measurement of the appearance oscillation channel is not possible (due to the missing capability to identify charges), there is good sensitivity to the sub-leading oscillation parameters $`\mathrm{sin}^22\theta _{13}`$ and the sign of the mass squared difference. Taking into account only statistical errors, we calculated that a measurement of $`\theta _{13}`$ would be possible down to $`\mathrm{sin}^22\theta _{13}`$ values of $`210^3`$ . In case of a negative sign of $`\mathrm{\Delta }m_{31}^2`$ the sensitivity would be worse, but this could be overcome by using an antineutrino beam. The values given above were all obtained at a baseline of 6500 km. With larger baselines, a test of the MSW-effect would be possible. In particular, at 11200 km the the sign of $`\mathrm{\Delta }m_{31}^2`$ (which can be revealed only through matter effects) can be determined down to $`\mathrm{sin}^22\theta _{13}`$ values of approximately $`10^2`$, depending on the precise value of $`|\mathrm{\Delta }m_{31}^2|`$.
An important aspect of the experimental scenario studied in this work is that a test of the MSW effect and the determination of $`\mathrm{sin}^22\theta _{13}`$ could be done with existing technologies. This type of high rate neutrino experiment might thus be an interesting alternative to neutrino factory experiments. Further studies of systematic errors in the beam flux and backgrounds are however necessary and we recommend that detailed simulations of the detector response should be performed. A study of the potential of $`\nu `$-factory beams pointing to a km<sup>3</sup> Cherenkov detector will be published soon .
Acknowledgments: We thank M. Leuthold and C. Spiering for supplying us with crucial information on detector issues. Furthermore, we want to thank L. Oberauer for helpful discussions. This work was supported by the โSonderforschungsbereich 375 fรผr Astro-Teilchenphysikโ der Deutschen Forschungsgemeinschaft.
|
warning/0006/cond-mat0006360.html
|
ar5iv
|
text
|
# References
Critical points in two-dimensional replica sigma models
Paul Fendley
Department of Physics
University of Virginia
Charlottesville VA 22901, USA
fendley@virginia.edu
Abstract
I survey the kinds of critical behavior believed to be exhibited in two-dimensional disordered systems. I review the different replica sigma models used to describe the low-energy physics, and discuss how critical points appear because of WZW and theta terms.
The last few years have seen a remarkable resurgence in activity on disordered systems in two dimensions. Even though serious study in this field goes back more than twenty years, recently there have been a number of precise quantitative results. The most famous experimental problem of this sort is the transition between plateaus in the quantum Hall effect. The experiments suggest strongly that there is a critical point in between the plateaus, and all the theoretical explanations of the quantum Hall effect indicate that disorder plays a crucial role in this phase transition.
In this talk, I will discuss the kinds of critical behavior which happen in systems of electrons which interact only with disorder. It is possible that electron interactions will substantially affect the problem. However, until we theoretically understand the problem without interactions, it is of course unlikely that we will be able to understand the role of electron interactions.
In Hamiltonians bilinear in electron annihilation/creation operators are classified. The disorder appears in the couplings of the Hamiltonian, i.e. the coefficients of the various terms are taken to be random with some distribution. For a system with a finite number of states, each Hamiltonian is some matrix belonging to the appropriate symmetry class (e.g. Hermitian, real symmetric, etc.). The simplest (zero-dimensional) way to analyze these Hamiltonians would be to consider an ensemble of random matrices belonging to the appropriate symmetry class. It turns out these classes are related to symmetric spaces, so the random matrix ensembles are conveniently labelled by a corresponding symmetric space . I present these labels in the first table, with a brief physical description of the corresponding systems. For example, models in the first three classes have a Hamiltonian describing electrons with spin hopping on a lattice, with or without $`SU(2)`$ spin symmetry $`๐ฎ`$ and a discrete time-reversal symmetry $`๐ฏ`$. Classes $`A`$III and $`C`$II describe electrons with no spin, but their Hamiltonians allow Cooper pairing terms which simultaneously annihilate or create a pair of electrons. The final four Hamiltonians have spin and Cooper pairing terms. The descriptions in parentheses will be discussed later in this talk.
| RMT | description |
| --- | --- |
| $`A`$ (GUE) | Anderson localization, broken $`๐ฎ,๐ฏ`$ (Hall plateau) |
| $`A`$I (GSE) | Anderson localization, broken $`๐ฎ`$, good $`๐ฏ`$ |
| $`A`$II (GOE) | Anderson localization, good $`๐ฎ,๐ฏ`$ |
| $`A`$III | spinless superconductor, broken $`๐ฏ`$ |
| $`BD`$I | |
| $`C`$II | spinless superconductor, good $`๐ฏ`$ |
| $`C`$ | superconductor, broken $`๐ฏ`$, good $`๐ฎ`$ (SQHE, $`d_{x^2y^2}+id_{xy}`$) |
| $`C`$I | superconductor, good $`๐ฏ,๐ฎ`$ ($`d_{x^2y^2}`$) |
| $`D`$ | superconductor, broken $`๐ฏ,๐ฎ`$ (random-bond Ising) |
| $`D`$III | superconductor, good $`๐ฏ`$, broken $`๐ฎ`$ |
Table 1: Random matrix theories and some physical descriptions
It is worth noting that properties at non-zero frequency $`\omega 0`$ are described by the GOE, GSE or GUE universality classes; the other universality classes only describe $`\omega 0`$ properties.
Since here I am interested in these systems in two spatial dimensions, I will not utilize random matrix theory. However, because classification according to random matrix ensembles is very convenient for these sorts of Hamiltonians, most workers in this field still refer to the systems by the names in the first column of table 1.
The disordered systems in table I can be described by certain sigma models in the replica formulation. I review here what these words mean.
Disorder is introduced into these fermion systems by allowing their couplings to vary randomly with some distribution. The action is $`S(\psi ,R)`$, where the fermions are denoted $`\psi `$ and the random variables $`R`$. The type of disorder considered here is called quenched, meaning physical quantity is computed first for fixed disorder, and then the randomness is averaged over. For example, the free energy at fixed disorder $`\mathrm{ln}Z(R)`$ follows from the path integral over the fermions for a given configuration $`R`$:
$$\mathrm{ln}Z(R)=\mathrm{ln}\left[[๐\psi ]e^{S(\psi ,R)}\right].$$
The physical free energy $`\overline{f}`$ is then given by averaging over the random variables $`R`$:
$$\overline{f}=[๐R]\mathrm{ln}Z(R).$$
(1)
The replica trick is a method for computing physical quantities which seems to work at least some of the time. For the free energy, the replica trick uses the identity
$$\mathrm{ln}Z=\text{lim}_{N0}\frac{Z^N1}{N}$$
to rewrite $`\mathrm{ln}Z(R)`$ in (1). The partition function to the $`N^{\text{th}}`$ power can be written as the product of $`N`$ path integrals over $`N`$ โreplicaโ fields $`\psi _\mu `$ where $`\mu =1\mathrm{}N`$, but all with the same random fields $`R`$. Then one makes the bold assumption that the $`N0`$ limit and the $`N+1`$ path integrals commute, giving
$$\overline{f}=\text{lim}_{N0}\frac{1}{N}\left[1+[๐R][๐\psi _1]e^{S(\psi _1,R)}[๐\psi _2]e^{S(\psi _2,R)}\mathrm{}[๐\psi _N]e^{S(\psi _N,R)}\right].$$
This trick thus puts the random fields on the same footing as the replica fields, and one can use standard field theory techniques on the problem.
The catch is that at the end of the computation one needs to set the number of fields $`N`$ to be zero. Since rarely in a field theory does one understand the analyticity properties of $`N`$, this step is certainly not rigorous, or even close. Below I will take the strategy of just doing computations at positive integer $`N`$ where everything is well defined, and hope that the formulas are still valid when $`N`$ is set to be zero. This procedure is well defined perturbatively, but we will attempt to extract non-perturbative information as well. There is another approach to these problems which utilizes supergroup symmetries instead of the replica trick. This approach does not suffer from all the possible ambiguities in taking this $`N0`$ limit. However, it is not known how to utilize the techniques of integrability in the supergroup formulation, so I am stuck with the replica trick.
To make further progress on the problem we map the problem onto a sigma model . The sigma model is an effective field theory which follows (uniquely) by analyzing the symmetries of the replica field theory. A sigma model is a field theory where the fields take values on some manifold. In this talk, all of these manifolds are of the form $`G/H`$, where $`G`$ and $`H`$ are Lie groups, with $`H`$ a maximal subgroup of $`G`$. Such a manifold is called a symmetric space. The best-known example of a sigma model on a symmetric space is often called the sphere sigma model, where $`G=O(3)`$ and $`H=O(2)`$. The field can be thought of as a vector with three components and fixed length, hence a sphere. While the vector can be rotated by the $`O(3)`$ group, it is invariant under the $`O(2)`$ subgroup consisting of rotations around its axis. Thus the space of distinct three-dimensional fixed-length vectors (the two-sphere) is the coset $`O(3)/O(2)`$.
The replica sigma models describing the models in Table 1 can read off the tables of . To see where they come from, it is useful to do one example in detail. This discussion I steal from . The Hamiltonian describes spinless fermions with a triplet p-wave type pairing:
$$H_0=\underset{k}{}ฯต_kc_k^{}c_k+(\mathrm{\Delta }_kc_k^{}c_k^{}+\mathrm{h}.\mathrm{c}.)$$
(2)
where $`\mathrm{\Delta }_k\frac{v_\mathrm{\Delta }}{2}k_x`$. Disorder is weak enough to maintain some notion of the Fermi surface. The low energy excitations of the fermions are then found about two nodes positioned on the $`k_y`$-axis at $`K_\pm =(0,\pm K)`$. Linearizing the theory about these nodes via $`ฯต_{K_\pm +q}=q_yv_F`$ and $`cc_1\mathrm{exp}(iK_+x)+c_2\mathrm{exp}(iK_{}x)`$, $`H`$ becomes
$$H=d^2x\psi ^{}(iv_F\tau _z_y+iv_\mathrm{\Delta }\tau _x_x)\psi .$$
(3)
where $`\psi ^{}=(c_1^{},c_2)`$. The Pauli matrices $`\tau _i`$ act in the particle-hole space of the spinors. $`H`$ is precisely the Hamiltonian of a single Dirac fermion in $`2+1`$ dimensions. One can compute correlators at fixed frequency $`\omega `$ by utilizing the action of a classical two-dimensional Euclidean field theory:
$$S_0=d^2x\psi ^{}(iv_F\tau _z_y+iv_\mathrm{\Delta }\tau _x_xi\omega \tau _z)\psi .$$
(4)
One way of adding randomness is to include the term
$$H_{\mathrm{disorder}}=R(c_1^{}c_1+c_2^{}c_2)$$
(5)
in the Hamiltonian. Here $`R`$ is a random variable with variance $`R(x)R(y)=u^1\delta (xy)`$; because of the delta function the disorder is called on-site. After replicating the fermions, $`\psi \psi _k`$, the path integral over the random field $`R`$ is easily done. This couples the different replicas via the quartic terms $`S_{\mathrm{disorder}}=\frac{1}{u}(\psi _k^{}\tau ^z\psi _k)(\psi _l^{}\tau ^z\psi _l)`$. Reorganizing the fields via $`\stackrel{~}{\psi }^{}(\psi _R^{},\psi _L^{})=\psi ^{}\tau ^z\mathrm{exp}(i\pi \tau ^x/4)`$ and $`\stackrel{~}{\psi }(\psi _R,\psi _L)=\mathrm{exp}(i\pi \tau ^x/4)\psi `$ gives the action
$$S=d^2xv_F\stackrel{~}{\psi }_k^{}(i_y\tau ^z_x)\stackrel{~}{\psi }_k\frac{1}{u}(\stackrel{~}{\psi }_k^{}\stackrel{~}{\psi }_k)(\stackrel{~}{\psi }_l^{}\stackrel{~}{\psi }_l).$$
(6)
This theory is invariant under the group $`U(N)_L\times U(N)_R`$ transforming $`\stackrel{~}{\psi }\frac{1}{2}((1+\tau ^z)U_L+(1\tau ^z)U_R)\stackrel{~}{\psi }`$. Adding disorder to the Cooper-pairing term results in another four-fermion term, but preserves this symmetry. As long as the symmetry structure is unchanged, the low-energy physics should be the same.
The sigma model describes the low-energy behavior of this system. The idea is familiar from many different contexts in condensed-matter and particle physics. The theory has a global symmetry $`G`$, which for this example is $`U(N)\times U(N)`$. However, we assume there is an energy scale where some fermion bilinear gets an expectation value. This expectation value is invariant under only a subgroup $`H`$ of the full symmetry group $`G`$. In dimensions higher than two, this would result in spontaneous symmetry breaking. The sigma model describes the interactions of the resulting Goldstone bosons, which take values on the space $`G/H`$. In two dimensions, the symmetry does not break spontaneously ($`G`$ remains a good global symmetry of the low-energy theory), but still the low-energy degrees of freedom live on $`G/H`$. The easiest way to see this explicitly is to introduce a Hubbard-Stratonovich matrix field $`M_{kl}`$ to factor the four-fermion term:
$$S=d^2x\left[iv_F\stackrel{~}{\psi }_k^{}(_y\tau ^z_x)\stackrel{~}{\psi }_k\frac{1}{u}(\stackrel{~}{\psi }_k^{}M_{kl}\stackrel{~}{\psi }_l)+\text{tr}M^2\right]$$
(7)
where $`M`$ is hermitian. Under the symmetry $`U(N)_L\times U(N)_R`$, $`MUMU^{}`$. The sigma model describes the physics around saddle points of this path integral. For example, one can have saddle points where $`M`$ is off-diagonal, e.g. $`M_{LL}=M_{RR}=0`$, but $`M_{LR}=M_{RL}^{}I`$, the identity. The diagonal subgroup $`U(N)_V`$ leaves this saddle point invariant, and so the low-energy modes $`T=M_{RL}`$ take values in $`U(N)_L\times U(N)_R/U(N)_VU(N)`$. Focusing solely upon these modes gives
$$S=S_0\frac{1}{u}(\stackrel{~}{\psi }_R^{}T\stackrel{~}{\psi }_L+\stackrel{~}{\psi }_L^{}T^{}\stackrel{~}{\psi }_R).$$
(8)
Integrating out the fermions leaves an effective action for the bosonic field $`T`$, which can be expanded in in powers of the momentum over the expectation value of $`T`$. I will discuss the form of this action later. The important thing is that it follows solely from the symmetry.
This model is in class $`A`$III in Table 1. This result can be read off from the tables in . The replica sigma model corresponding to a given disordered universality class is determined by taking $`G/H`$ to be the bosonic subspace labeled โ$`M_F`$โ in the second table in . The $`F`$ is for fermion; we are using fermionic replicas here. If we instead had used bosonic replicas to compute the same Greenโs functions, we would have ended up with a replica sigma model on a non-compact space; this model is labelled $`M_B`$ in Zirnbauerโs table. For the above example, this would be $`Gl(N,C)/U(N)`$. The non-compact sigma model is not equivalent to the compact one except hopefully in the replica limit $`N0`$. However, very few exact results are available for sigma models on non-compact spaces (in fact, the difficulty in obtaining exact results for the supergroup sigma models arises mainly from the non-compact bosonic subgroup, not from the fermionic fields).
I will concentrate here exclusively on the replica sigma models on compact spaces. I present a two-dimensional version of Zirnbauerโs tables in Table 2. The first column is the commonly-used label coming from figure 1; I have rearranged the rows for reasons which will become clear later. The second column is the replica sigma model. The third column lists the types of critical behavior possible in this sigma model. It is the purpose of the rest of this paper to discuss the third column.
| RMT | replica sigma model | possible 2D critical behavior |
| --- | --- | --- |
| GUE | $`U(2N)/U(N)\times U(N)`$ | Pruisken phase |
| $`C`$ | $`Sp(2N)/U(N)`$ | Pruisken phase |
| $`D`$ | $`O(2N)/U(N)`$ | Pruisken phase, metallic phase |
| $`C`$II | $`U(N)/O(N)`$ | $`\theta =\pi U(N)_1`$; Gade phase |
| GSE | $`O(2N)/O(N)\times O(N)`$ | $`\theta =\pi O(2N)_1`$; metallic phase |
| $`A`$III | $`U(N)\times U(N)/U(N)`$ | WZW term; Gade phase |
| $`C`$I | $`Sp(2N)\times Sp(2N)/Sp(2N)`$ | WZW term |
| $`D`$III | $`O(N)\times O(N)/O(N)`$ | WZW term; metallic phase |
| $`BD`$I | $`U(2N)/Sp(2N)`$ | Gade phase? |
| GOE | $`Sp(4N)/Sp(2N)\times Sp(2N)`$ | none! |
Table 2: The replica sigma models and their possible critical behavior
The sphere sigma model we have discussed is the $`N=1`$ case of the classes GUE, $`C`$, $`C`$II, and the $`N`$=2 case of classes $`D`$ and GSE (the latter comprises two copies of the sphere).
It is important to note that even though each random matrix theory is associated with a symmetric space as discussed in (this is the origin of most of the names), this symmetric space is not necessarily the same as the symmetric space of the corresponding replica sigma model. For example, the random matrices for the theory $`C`$I are in the tangent space of $`Sp(2N)/U(N)`$, (i.e. exponentiating the random matrices gives the space $`C`$I$`Sp(2N)/U(N)`$), but the corresponding replica sigma model is on the symmetric space $`Sp(2N)\times Sp(2N)/Sp(2N)`$.
Sigma models on symmetric spaces have the convenient property that they have only one coupling constant. Roughly speaking, this means that they preserve their โshapeโ under renormalization: only their size changes. For example, the sphere sigma model remains a sphere under renormalization: only the radius of the sphere renormalizes. Actually, the sigma models for $`A`$III, $`C`$II and $`BD`$I are not quite symmetric spaces: the corresponding symmetric spaces have $`SU(N)`$ instead of $`U(N)`$. The extra $`U(1)`$ has some interesting effects (for example turning a critical point into a critical line), but does not affect the sigma model on the symmetric space at all.
At low energy, one can safely neglect four-derivative terms and higher in the action. All the above โordinaryโ sigma models have an action can be written in the form
$$S_{ordinary}=g๐x๐y\text{tr}\left[_\mu T^\mu T^1\right]$$
(9)
where $`T`$ is a unitary matrix, possibly with further restrictions. This action is very non-trivial because of the non-linear constraint of unitarity. For the example above, $`T`$ is an $`N\times N`$ unitary matrix, because $`U(N)\times U(N)/U(N)U(N)`$ as a space. For $`T`$ in $`U(N)/O(N)`$ (class $`C`$II), $`T`$ is an $`N\times N`$ symmetric unitary matrix. Symmetric matrices do not close under multiplication, which is why $`U(N)/O(N)`$ is not a group but rather a coset. For $`O(2N)/O(N)\times O(N)`$ (class GSE), $`T`$ is a $`2N\times 2N`$ symmetric real orthogonal matrix.
An important observation of Andersonโs is that the coupling constant $`g`$ of the sigma model is related to the conductance of the system: if $`g`$ renormalizes to zero, the system is localized and not conducting, while if $`g`$ renormalizes to infinity, the system is metallic. The latter is the trivial fixed point of the system. The latter corresponds to the size of the manifold (e.g. the radius of the sphere) going to infinity , which means the manifold becomes effectively flat. For ordinary sigma models on compact symmetric spaces with $`N>1`$, the trivial fixed point is unstable. The beta function is proportional to the curvature of the manifold , which is always positive for these spaces. Moreover (for $`N>1`$), the beta function shows no evidence for a non-trivial fixed point. This is a more or less a consequence of the Mermin-Wagner-Coleman theorem, which says that continuous symmetries can not be spontaneously broken in two dimensions. This implies that the manifold should be always curved (and not critical); otherwise there would be Goldstone modes in violation of the theorem.
However, in two dimensions, there are a number of different ways critical behavior appears in a disordered system. I group them into three types:
* 1. Perturbative Peculiarities
a) Gade phase
b) Metallic phase
* 2. WZW term
* 3. Theta term
a) Pruisken type (non-vanishing $`\sigma _{xy}`$)
b) $`\theta =0`$ or $`\pi `$ only
We will discuss all these methods, but devote most of our attention to 2 and 3. These require adding extra terms to the action $`S_{ordinary}`$.
1. Perturbative Peculiarities
While it is important to note that the Mermin-Wagner-Coleman theorem implies no non-trivial fixed points in ordinary sigma models for $`N>1`$, this of course does not prohibit interesting things from happening in the replica limit $`N0`$ . For example, the dimension of all the manifolds in table II goes to zero as $`N0`$, making the idea of positive curvature somewhat confusing. In fact, a while ago (see and references within) it was shown that at one loop
$`\beta `$ $`N`$ $`\text{if }G=U(N)`$
$`\beta `$ $`N2`$ $`\text{if }G=O(N)`$
$`\beta `$ $`N+1`$ $`\text{if }G=Sp(2N)`$
where the constant of proportionality is a negative number for $`N0`$. Therefore, by โperturbative peculiaritiesโ I mean the consequences of the fact that as $`N0`$ the $`\beta `$ function goes to zero for $`G=U(N)`$ and has changed sign for $`G=O(N)`$.
If the beta function goes to zero to all orders as $`N0`$, this opens up the possibility that there is no flow in the sigma model: for any $`g`$ the model is at a fixed point. This indeed happens for classes $`A`$III and $`C`$II (and probably $`BD`$I, although I am not aware of any explicit computations other than of the perturbative beta function). This behavior was discussed at length in , which is why I call this the Gade phase. In Gadeโs work these universality classes are realized by a particle hopping on a bipartite lattice (the particle is restricted to hop only from one sublattice to the other); class $`C`$II has time-reversal symmetry, while class $`A`$III does not. The supergroup approach to these models was discussed in detail in . Since these models also have an extra $`U(1)`$ factor as mentioned above, the model is critical over an entire plane of couplings.
If the beta function is positive as $`N0`$, the trivial fixed point is stable. This happens for classes $`D`$, $`D`$III and GSE. Since $`g`$ renormalizes to infinity, this phase is conducting and is hence called metallic. This implies the existence of at least one non-trivial fixed point, because it is still expected that for strong enough disorder, the system does not conduct. Hence there should be a phase transition from the metallic phase to a localized phase at some value of $`g`$. This fixed point should be unstable in $`g`$ in both directions (i.e. for $`g<g_c`$ the system renormalizes to $`g0`$, while for $`g>g_c`$ the system renormalizes to $`g\mathrm{}`$).
2. WZW term
Two-dimensional sigma models with a manifold of the form $`H\times H/H`$ are called principal chiral models, and have been widely studied. They are massive asymptotically-free field theories for $`N>1`$. However, there is an additional term which can be added to the action (9) which changes the low-energy behavior from gapped to gapless for any $`N`$. This is called the Wess-Zumino-Witten term. To write it out explicitly, first one needs to consider field configurations $`h(x,y)`$ which fall off at spatial infinity, so that one can take the spatial coordinates $`x`$ and $`y`$ to be on a sphere. Then one needs to extend the fields $`h(x,y)`$ on the sphere to fields $`h(x,y,z)`$ on a ball which has the original sphere as a boundary. The fields inside the ball are defined so that $`h`$ at the origin is the identity matrix, while $`h`$ on the boundary is the original $`h(x,y)`$. It is possible to find a continuous deformation of $`h(x,y)`$ to the identity because $`\pi _2(H)=0`$ for any simple Lie group. Then the WZW term is $`k\mathrm{\Gamma }`$, where
$$\mathrm{\Gamma }=\frac{ฯต_{abc}}{24\pi }๐x๐y๐z\text{tr}\left[(h^1_ah)(h^1_bh)(h^1_ch)\right].$$
(10)
The coefficient $`k`$ is known as the level, and for compact groups must be an integer because the different possible extensions of $`h(x,y)`$ to the ball yield a possible ambiguity of $`2n\pi `$ in $`\mathrm{\Gamma }`$. The WZW term changes the equations of motion and beta function, but only by terms involving $`h(x,y)`$: the variation of the integrand is a total derivative in $`z`$.
The two-dimensional sigma model with WZW term has a stable fixed point at $`1/g=16\pi /k`$ , so the model is critical and the quasiparticles are gapless. The corresponding conformal field theory is known as the $`H_k`$ WZW model . The WZW term is invariant under discrete parity transformations (e.g. $`xx`$, $`yy`$) if $`hh^1`$ under this transformation. For there to be a WZW term in a parity-invariant theory, some of the low-energy fields must be pseudoscalars.
The WZW term allows some of these sigma models which has a stable low-energy fixed point. Why this term must arise in many situations was understood in particle physics some time ago. This was the topic of . The WZW term was shown to arise in several disordered systems , by bosonizing the explicit system. There it appears to ensure the certain current-algebra commutation relations are obeyed properly in the bosonic theory. It appears more generally in the map of the action like (7) to a sigma model. The sigma model arises when integrating out the fermions in this action, leaving a low-energy effective action for $`h`$. Upon doing so, one easily obtains the ordinary sigma model action of the form (9) (with $`T=h`$). The WZW term arises for a subtler reason. To perform a consistent low-energy expansion of the action , one must change field variables. This results in a Jacobian in the path integral , which in these two-dimensional cases is precisely the WZW term (10).
In fact, one can determine without these explicit computations whether or not the WZW term will appear in the low-energy effective action. The reason is some very deep physics known as the chiral anomaly. In the models which admit a WZW term, the fermions have a chiral $`H_L\times H_R`$ symmetry. As is well known, chiral symmetries involving fermions are frequently anomalous. Noetherโs theorem says that a symmetry of the action gives a conserved current with $`_\mu j^\mu =0`$, but this is only true to lowest order in perturbation theory. An anomaly is when the current is not conserved in the full theory (although the associated charge is still conserved). In the case of massless fermions in $`1+1`$ dimensions, this was shown in detail in . The WZW term is the effect of the anomaly on the low-energy theory. Even though the fermions effectively become massive when $`T`$ gets an expectation value, their presence still has an effect on the low-energy theory, even if this mass is arbitrarily large. This violation of decoupling happens because the chiral anomaly must be present in the low-energy theory. In other words, the anomaly coefficient does not renormalize. This follows from an argument known as โt Hooft anomaly matching . One imagines weakly gauging the anomalous symmetry. It is not possible to gauge an anomalous symmetry in a renormalizable theory, but one can add otherwise non-interacting massless chiral fermions to cancel the anomaly. Adding these spectator fermions ensures that the appropriate Ward identities are obeyed, and the symmetry can be gauged. In the low-energy effective theory, the Ward identities must still be obeyed. Because the massless spectator fermions are still present in the low-energy theory, there must be a term in the low-energy action which cancels the anomaly from the spectators. This is the WZW term.
To determine whether a chiral anomaly and hence a WZW term is present, one usually needs to do only a simple one-loop computation. For chiral symmetries, it is customary to define the vector and axial currents $`j_V^\mu =j_L^\mu +j_R^\mu `$, $`j_A^\mu =j_L^\mu j_R^\mu `$; for these theories $`j_A^\mu =ฯต^{\mu \nu }j_{V\nu }`$. One then computes the correlation functions $`j_V^\mu (x,y)j_V^\nu (0,0)C^{\mu \nu }`$. If $`_\mu C^{\mu \nu }0`$ or $`ฯต_{\alpha \beta }^\alpha C^{\beta \nu }0`$ for some $`\nu `$, then there is an anomaly. An important characteristic of the anomaly is that it is independent of any continuous change in the theory, as long as the chiral symmetry is not broken explicitly. Thus to find the anomaly, one can compute $`C^{\mu \nu }`$ using free fermions (where the only contribution is a simple one-loop graph, see e.g. ).
For the $`p`$-wave superconductor example discussed in detail above (class $`A`$III), one has $`H=U(N)`$ and the possibility of a WZW term. With two nodes in the fermi surface as discussed above, one in fact does, with $`k=1`$ . The model is already critical because it is in a Gade phase, but the WZW term does have an effect; for example when the coupling $`g`$ is at the WZW fixed point, the density of states falls off with a power law in energy instead of an exponential .
Classes $`C`$I and $`D`$III are more interesting in this context.
Class $`C`$I basically amounts to generalizing (2) to include the spin of the electron, and preserving the $`SU(2)`$ spin symmetry . As discussed in , if there are two nodes in the fermi surface, one ends up with an $`Sp(2N)_1=U(N)_2`$ critical point. The case of most interest is the $`d_{x^2y^2}`$ superconductor, where there are four nodes in the Fermi surface. If the two sets of nodes are not coupled then the global symmetry is enlarged to $`Sp(2N)\times Sp(2N)\times Sp(2N)\times Sp(2N)`$. โt Hooftโs argument means that one obtains two copies of the $`Sp(2N)_1`$ theory with conserved currents
$$_\mu j_{V1}^\mu =_\mu j_{V2}^\mu =0$$
$$_\mu j_{A1}^\mu =_\mu j_{A2}^\mu 0.$$
The reason the anomalies in the second line are the same is that I have defined the two sets of fields with the same conventions (i.e. the same sign WZW term). However, generic disorder couples all the nodes, breaking the symmetry back to $`Sp(2N)\times Sp(2N)`$. If the remaining symmetries are generated by $`j_{V1}^\mu +j_{V2}^\mu `$ and $`j_{A1}^\mu +j_{A2}^\mu `$, then the latter is anomalous. Because of โt Hooftโs argument, the two anomalies just add and one obtains $`Sp(2N)_2`$ . However, the conserved symmetries of the $`d_{x^2y^2}`$ superconductor are determined by the band structure, and this requires that the conserved currents with these conventions be $`j_{V1}^\mu +j_{V2}^\mu `$ and $`j_{A1}^\mu j_{A2}^\mu `$ . The anomalies in the latter cancel, and therefore there is no WZW term and no phase transition in the resulting sigma model. I have used relativistic notation in this section, but even if 1+1 dimensional Lorentz invariance is broken (e.g. by taking $`v_{F1}/v_{\mathrm{\Delta }1}v_{F2}/v_{\mathrm{\Delta }2}`$), these results still apply.
A WZW term likewise appears in class $`D`$III when there are two nodes. For $`N>2`$ this fixed point is stable like the others. As discussed above, the sign of the beta function has flipped as $`N0`$, so the model has a metallic phase. This makes the $`N0`$ limit of the $`O(2N)_k`$ fixed point an excellent candidate for the unstable fixed point between the metallic and localized phases. The subtlety here is that it is not clear whether or not one can continue the well-understood $`N>2`$ results to the replica limit. There is at least one well-known situation (two-dimensional self-avoiding polymers) where one cannot do such a continuation. I am currently studying this question, but do not yet have any conclusive answers.
There are two morals of this section:
* 1. When the WZW term is present, the model has a non-trivial fixed point.
* 2. Whether or not the WZW term is present depends on original (microscopic) disordered model considered. For the case of fermions, the coefficient of this term is easily determined.
3. $`\theta `$ term
Another term which can be added to some sigma model actions is called the theta term. This term is inherently non-perturbative: it does not change the beta function derived near the trivial fixed point at all. Nevertheless, it can result in a non-trivial fixed point.
The theta term is best illustrated in the sphere sigma model. As above, I consider field configurations which go to a constant at spatial infinity, so the spatial coordinates are effectively that of a sphere. Since the field takes values on a sphere, the field is therefore a map from the sphere to a sphere. An important characteristic of such maps is that they can have non-trivial topology: they cannot necessarily be continuously deformed to the identity map. This is analogous to what happens when a circle is mapped to a circle (i.e. a rubber band wrapped around a pole): you can do this an integer number of times called the winding number (a negative winding number corresponds to flipping the rubber band upside down). It is the same thing for a sphere: a sphere can be wrapped around a sphere an integer number of times. An example of winding number 1 is the isomorphism from a point on the spatial sphere to the same point on the field sphere. The identity map is winding number 0: it is the map from every point on the spatial sphere to a single point on the field sphere, i.e. $`T(x,y)=`$constant. In field theory, field configurations with non-zero winding number are usually called instantons. The name comes from viewing one of the directions as time (in our case, one would think of say $`x`$ as space and $`y`$ as Euclidean time). Since instanton configurations fall off to a constant at $`y=\pm \mathrm{}`$, the instanton describes a process local in time and hence โinstantโ. In an even fancier modern language, an instanton would be called a $`1`$ brane. Therefore, the field configurations in the sphere sigma model can be classified by an integer $`n`$. We can therefore add a term
$$S_\theta =in\theta $$
to the action, where $`\theta `$ is an arbitrary parameter. Since $`n`$ is an integer, the physics is periodic under shifts of $`2\pi `$ in $`\theta `$.
Pruisken showed that the replica sigma model describing two-dimensional non-interacting electrons with disorder and a strong transverse magnetic field (which breaks $`๐ฏ`$) is in the GUE class. This model has instantons with integer winding number, and so allows a $`\theta `$ term. He showed that while the sigma model coupling $`g`$ is related to the conductivity $`\sigma _{xx}`$, the other parameter $`\theta `$ is related to the Hall conductivity $`\sigma _{xy}`$. He proposed that at $`\theta =\pi `$, the system has a critical point separating a phase with $`\sigma _{xy}=0`$ from $`\sigma _{xy}=1`$: the famous (experimentally observed) transition between quantum Hall plateaus. This critical point is stable in $`g`$ but unstable when $`\theta `$ is taken away from $`\pi `$. While Pruiskenโs proposed phase diagram is widely believed to be correct, noone has succeeded in deriving any analytic results valid for the replica limit $`N0`$. The best evidence that Pruiskenโs phase diagram does apply to GUE class models is indirect, coming from numerical studies of the network model . The network model can be mapped on to a supergroup spin chain whose continuum limit should be described by a supergroup sigma model of the GUE class. Thus ven though the network model is microscopically different from the model of electrons with disorder and transverse magnetic field, it should be in the same universality class. Numerical studies are much easier to do on the network model or on the supergroup spin chain, and the work done is completely consistent with this phase diagram.
Although we do not have any exact results applicable the GUE replica limit, we do a number of exact results for some sigma models at $`\theta =\pi `$, and I would like to discuss them in this section. These all are in harmony with Pruiskenโs picture. Although Pruisken did not know this at the time of his proposal (he was reasoning by analogy with some of โt Hooftโs work on QCD ), the sphere sigma model has the same structure as he proposed for the Hall plateaus. Another way of saying this would be to say the phase structure of the $`U(2N)/U(N)\times U(N)`$ sigma model is believed to be the same for $`N=1`$ and $`N0`$. While we do not know the exact nature of the non-trivial critical point for $`N0`$, it is understood well for the sphere, $`N=1`$. Namely, Haldane realized when studying the half-integer-spin Heisenberg spin chains that the sphere sigma model at $`\theta =\pi `$ has a non-trivial fixed point stable in $`g`$. This fixed point turns out to be exactly the $`SU(2)_1`$ WZW model . The argument goes as follows. First one uses Zamolodchikovโs c-theorem, which makes precise the notion that as one follows renormalization group flows, the number of degrees of freedom goes down. Zamolodchikov shows that there is a quantity $`c`$ associated with any two-dimensional unitary field theory such that $`c`$ must not increase along a flow. At a critical point, $`c`$ is the central charge of the corresponding conformal field theory . At the trivial fixed point of a sigma model where the manifold is flat, the central charge is the number of coordinates of the manifold (i.e. the number of free bosons which appear in the action (9)). For the sphere, this means that $`c=2`$ at the trivial fixed point. The only unitary conformal field theories with $`SU(2)`$ symmetry and $`c<2`$ are $`SU(2)_k`$ for $`k<4`$ (in general, the central charge of $`SU(2)_k`$ is $`3k/(k+2)`$). One can use the techniques of to show that there are relevant operators at these fixed points, and at $`k=2`$ or $`3`$, no symmetry of the sphere sigma model prevents these relevant operators from being added to the action . So while it is conceivable that the sphere sigma model with $`\theta =\pi `$ could flow near to these fixed points, these relevant operators would presumably appear in the action and cause a flow away. However, there is only one relevant operator (or more precisely, a multiplet corresponding to the field $`h`$ itself) for the $`SU(2)_1`$ theory. The sigma model has a discrete symmetry $`TT`$ when $`\theta =\pi `$; the winding number $`n`$ goes to $`n`$ under this symmetry, but $`\theta =\pi `$ and $`\theta =\pi `$ are equivalent because of the periodicity in $`\theta `$. This discrete symmetry of the sigma model presumably turns into the symmetry $`hh`$ of the WZW model. While the operator $`\text{tr}h`$ is $`SU(2)`$ invariant, it is not invariant under this discrete symmetry. Therefore, the only possible low-energy fixed point for the sphere sigma model at $`\theta =\pi `$ is $`SU(2)_1`$. A variety of arguments involving the spin chain suggest strongly that this does in fact happen .
This picture also shows what happens when $`\theta `$ is moved away from $`\pi `$. Here the discrete symmetry is broken but $`SU(2)`$ is preserved, so one adds $`\text{tr}h`$ to the $`SU(2)_1`$ action. This is relevant, and in fact is equivalent to the sine-Gordon model (with dimension 1/2 $`\mathrm{cos}`$ term). This is a massive field theory, with no non-trivial low-energy fixed point. The sphere sigma model reproduces exactly Pruiskenโs phase diagram.
The flow of the sphere sigma model at $`\theta =\pi `$ to the $`SU(2)_1`$ WZW model was essentially proven in . This result does not seem to be widely known, so I will review it here. The sphere sigma model is integrable at $`\theta =0`$, meaning that there are an infinite number of conserved currents which allow one to find exactly the spectrum of quasiparticles and their scattering matrix in the correspoding $`1+1`$ dimensional field theory. There is evidence that $`\theta =\pi `$ case is integrable also, so one can assume so and go on to find the quasiparticle $`S`$ matrix here as well. This is done in . They find that while the quasiparticles for $`\theta =0`$ are gapped and form a triplet under the $`SU(2)`$ symmetry, for $`\theta =\pi `$ they are gapless, and form $`SU(2)`$ doublets (left- and right-moving). This is a beautiful example of charge fractionalization: the field $`T`$ is a triplet under the $`SU(2)`$ symmetry, but when $`\theta =\pi `$ the excitations of the system are doublets. They compute the $`c`$ function, and find that at high energy $`c`$ indeed is $`2`$ as it should be at the trivial fixed point, while $`c=1`$ as it should be at the $`SU(2)_1`$ low-energy fixed point. As an even more detailed check, the free energy at zero temperature in the presence of a magnetic field was computed for both $`\theta =0`$ and $`\pi `$ . The results can be expanded in a series around the trivial fixed point. One can identify the ordinary pertubative contributions to this series, and finds that they are the same for $`\theta =0`$ and $`\pi `$, even though the particles and $`S`$ matrices are completely different. This is as it must be: instantons and hence the $`\theta `$ term are a boundary effect and hence cannot be seen in ordinary perturbation theory. One can also identify the non-perturbative contributions to these series, and see that they differ. Far away from the trivial fixed point, non-perturbative contributions can dominate, which is why $`\theta =0`$ has no non-trivial fixed point, while $`\theta =\pi `$ does.
The question now is if similar behavior is found for any other disordered universality classes in two dimensions. In the sigma model language, the question is if any other of the models in Table 2 have instantons and hence allow a $`\theta `$ term. This question has already been answered by mathematicians; for a review accessible to physicists, see . In mathematical language, the question is whether the second homotopy group $`\pi _2(G/H)`$ is non-trivial. The second homotopy group is just the group of winding numbers of maps from the sphere to $`G/H`$, so for the sphere it is the integers. The general answer is that $`\pi _2(G/H)`$ is the kernel of the embedding of $`\pi _1(H)`$ into $`\pi _1(G)`$, where $`\pi _1(H)`$ is the group of winding numbers for maps of the circle into $`H`$. We have seen already that $`\pi _1(H)`$ is the integers when $`H`$ is the circle $`=U(1)=SO(2)`$. The only simple Lie group $`H`$ for which $`\pi _1`$ is nonzero is $`SO(N)`$, where $`\pi _1(SO(N))=๐_2`$ for $`N3`$ and $`๐`$ for $`N`$=2. Thus there are models with integer winding number, some with just winding number $`0`$ or $`1`$, and some with no instantons at all. Integer winding number means that $`\theta `$ is continuous and periodic, while a winding number of $`0`$ or $`1`$ means that $`\theta `$ is just $`0`$ or $`\pi `$ (just think of $`\theta `$ as being the Fourier partner of $`n`$). The results are summarized in the last column of Table 2; the models with integer winding number are labelled as having a Pruisken phase, while those with $`๐_2`$ winding number are $`C`$II and GSE.
The replica sigma models with integer winding number and continuous $`\theta `$ are believed to behave like the sphere sigma model. In addition to the GUE class, this happens for the $`Sp(2N)/U(N)`$ sigma models (class $`C`$) and the $`O(2N)/U(N)`$ sigma models (class $`D`$). The replica limit of class $`C`$ should have Pruiskenโs phase diagram, while in class $`D`$ it should be modified because of the flip in sign of the beta function: the non-trivial fixed point at $`\theta =\pi `$ should be unstable in $`\theta `$ and $`g`$, and another non-trivial fixed point should appear at some value $`g_c,\theta =0`$ (because the metallic phase should not exist at small enough $`g`$, i.e. strong enough disorder).
In all three of these universality classes, there is a network model (roughly speaking, one for each type of simple Lie group, $`U(N)`$, $`Sp(2N)`$ and $`O(2N)`$) . In all three cases, numerical results on the network model are consistent with the existence of a non-trivial fixed point as Pruisken predicted. Class $`D`$ turns out to be a complicated story . It seems that the sigma model does not describe all the physics of this class: because of the existence of domain walls, the complete phase diagram involves more than the two parameters $`\theta `$ and $`g`$ of the sigma model and Pruiskenโs phase diagram . The two-dimensional random-bond Ising model belongs to this symmetry class, but is a subspace of this full space. All results support the existence of a non-trivial critical point (or actually, points), but very little is known about detailed properties. On the other hand, the class $`C`$ model, known as the spin quantum Hall effect (SQHE) , is better understood. There is a remarkable exact result mapping certain correlators at the non-trivial fixed point onto a known conformal field theory, that describing percolation . There are no exact results from the replica sigma model point of view, but I have a conjecture to which I will return below.
It is not yet known whether the sigma models in the GUE, $`C`$ or $`D`$ classes are integrable. I do have a variety of exact results for the models with $`๐_2`$ instantons, namely classes $`C`$II and GSE . Without a continuous $`\theta `$ parameter, there does not seem to be any $`\sigma _{xy}`$, so these models do not have the full structure of the above three classes. However, these two models still have a non-trivial fixed point when $`\theta =\pi `$, and for this reason I believe they provide strong support for Pruiskenโs picture.
Basically, my results generalize those of to this much more general case. This is important because to have any hope of being able to take the replica limit, one needs a solution for any $`N`$. For class $`C`$II, the sigma models are on the space $`U(N)/O(N)`$. This sigma model has action (9) with $`T`$ a symmetric, unitary matrix. I find that when $`\theta =0`$, these sigma models are the $`U(N)`$ generalization of the sphere sigma model. When $`\theta =0`$, the model has a gap, with the spectrum consisting of massive particles in the symmetric representation of $`SU(N)`$ (plus bound states in more general representations). When $`\theta =\pi `$, the spectrum consists of gapless quasiparticles which are in the fundamental representations (vector, antisymmetric tensor,โฆ) of $`SU(N)`$. The non-trival low-energy fixed point when $`\theta =\pi `$ corresponds to $`SU(N)_1\times U(1)`$. Thus we see that at least for $`N>0`$, the replica sigma models in class $`C`$II with $`\theta =\pi `$ have exactly the same fixed point as those in class $`A`$III when a $`k=1`$ WZW term is present! The effect of having $`\theta =\pi `$ is as discussed before: the density of states changes its behavior near special values of the coupling.
The results for the GSE class are similar. This model is the $`O(2N)/O(N)\times O(N)`$ sigma model, which has action (9) with $`T`$ a symmetric, real and orthogonal matrix. When $`\theta =0`$, the model is gapped. When $`\theta =\pi `$, the model is gapless with a non-trivial stable fixed point corresponding to the $`O(2N)_1`$ WZW model. This sigma model proves to be the $`O(2N)`$ generalization of the sphere sigma model. The $`O(2N)_1`$ model turns out to be $`2N`$ free Majorana fermions. The word โfreeโ is slightly deceptive, because just as in the 2d Ising model, one can study correlators of the magnetization or โtwistโ operator, which are highly non-trivial. Because of the changes in the beta function at $`N`$=1, it is not clear yet whether these results can be continued to the replica limit; I am currently studying this. However, again it proves that the idea of a non-trivial critical point at $`\theta =\pi `$ is not a fluke of the sphere sigma model, and is true for any $`N`$.
The cases with $`๐_2`$ instantons are very similar to the WZW cases: $`\theta `$ is not a tunable parameter. In fact, I believe it is fixed uniquely by the underlying disordered system. Indeed, the expression for the winding number as an integral over the fields is precisely of the form of the WZW term. The deep connection between anomalies and theta terms was discussed in .
There are three models with a Pruisken phase, roughly corresponding to the three kinds of Lie groups $`U(N)`$, $`Sp(2N)`$ and $`O(2N)`$. There are only two models with $`๐_2`$ instantons, roughly corresponding to $`U(N)`$ and $`O(2N)`$ type, flowing to $`U(N)_1`$ and $`O(2N)_1`$ when $`\theta =\pi `$. It is logical to ask if there a sigma model with $`Sp(2N)`$ symmetry resembling the latter two. From the replica point of view, it is clearly the sigma model in class $`C`$, namely $`Sp(2N)/U(N)`$. The reason I view this as analogous is that the sigma model has action (9), where $`T`$ is unitary and symmetric like the other two, but with the additional restriction that tr$`(JT)=0`$, where $`J`$ is the $`2N\times 2N`$ matrix
$$J=\left(\begin{array}{cc}0& I\\ I& 0\end{array}\right),$$
where $`I`$ is the $`N\times N`$ identity matrix. One can presumably obtain the $`Sp(2N)/U(N)`$ sigma model from the $`SU(2N)/SO(2N)`$ model by pertubing by something like $`\lambda (\text{tr}JT)^2`$. This breaks the global $`SU(2N)`$ symmetry to $`Sp(2N)`$. The question is if when $`\theta =\pi `$, there remains a non-trivial critical point after perturbing. For $`N=1`$, the two sigma models are the same ($`Sp(2)=SU(2)`$) so obviously the fixed points are the same. For $`N>1`$, any non-trivial fixed point in $`Sp(2N)/U(N)`$ should be a perturbation of the $`\theta =\pi `$ fixed point of $`SU(2N)/SO(N)`$, namely $`SU(2N)_1`$, which has central charge $`2N2`$. If there is a non-trivial critical point of $`Sp(2N)/U(N)`$ when $`\theta =\pi `$, it must have central charge less than $`2N2`$, which leaves only $`Sp(2N)_1`$. Obviously, this is not a proof there is such a point, and moreover, it does not say if this behavior can be continued to $`N0`$. Nevertheless, if this correspondence holds, it predicts that in class $`C`$ the density of states $`\rho (E)E^{1/7}`$ . This agrees with the exact result of derived from the map onto percolation. It also predicts that there is another relevant operator of positive dimension $`5/4`$, which is thermal operator in percolation and the operator corresponding to moving off the critical point in the network model. These numbers also can be found from the analogous supersymmetric approach . While I think the agreement of dimensions is not a coincidence, this hardly proves that the non-trivial fixed point in class $`C`$ is $`Sp(2N)_1`$. It would be much more convincing if a correlator could be computed and shown to be equivalent correlator in percolation.
So this section has two morals virtually identical to those in the last:
* 1. All the available evidence suggests that when $`\theta =\pi `$, there is a non-trivial fixed point, in support of Pruiskenโs scenario.
* 2. For models with $`๐_2`$ instantons, the underlying disordered system should determine if $`\theta =\pi `$ or $`0`$.
I want to add a third moral:
* 3. Relevant operators may not always be relevant.
What I mean by the last is best illustrated by an example, following . Consider the principal chiral model on $`SU(2)`$ with the action (9), with the field $`T`$ taking values in $`SU(2)`$. Now add a $`k=1`$ WZW term, (10) with $`h=T`$. As noted before, this causes a flow to the stable fixed point $`SU(2)_1`$. Say in addition to adding the WZW term, I also add a term $`\lambda (\text{tr}T)^2`$. Around the trivial fixed point, this is a relevant perturbation, breaking the chiral symmetry but not the diagonal $`SU(2)`$. One might think it wrecks the flow to the $`SU(2)_1`$ (chirally-invariant) fixed point. However, it does not necessarily. An $`SU(2)`$ matrix $`T`$ can be rewritten as
$$\left(\begin{array}{cc}n_0+in_1& n_2+in_3\\ n_2+in_3& n_0in_1\end{array}\right)$$
where the otherwise-free parameters must satisfy $`(n_0)^2+(n_1)^2+(n_2)^2+(n_3)^2=1`$. The chiral-symmetry-breaking perturbation is $`\lambda (n_0)^2`$. For $`\lambda `$ large, its effect is to force $`n_0=0`$, leaving $`(n_1)^2+(n_2)^2+(n_3)^2=1`$. This is the sphere: this relevant pertubation turns the principal chiral model into the sphere sigma model. The WZW term turns into the theta term of the sigma model, with $`\theta =k\pi `$. The result discussed above shows that if $`k`$ is an odd integer, the presence of both the WZW term and the chiral-symmetry-breaking perturbation does not result in a massive theory: one ends up at the $`SU(2)_1`$ fixed point! For $`k=1`$, one ends up exactly where one would have otherwise, although the flow does reach the $`SU(2)_1`$ fixed point from a different (chirally non-invariant) direction. In fact one can see directly at the $`SU(2)_1`$ fixed point that all fields $`T_{\alpha \beta }T_{\gamma \delta }`$ operator is irrelevant there . However, I think this is a useful moral for the situation with an unknown low-energy fixed point: just because there is a relevant operator at the trivial fixed point doesnโt necessarily mean it will always be relevant.
I have discussed how a non-trivial fixed point can appear in a two-dimensional replica sigma model. These are summarized in Table 2. Every universality class save one has at least one kind of possible non-trivial critical point. Ironically, the only one that does not is Andersonโs original problem of free electrons with disorder!
I would like to thank Robert Konik for many conversations and for collaborating on . I have benefitted enormously from conversations and correspondence with John Chalker, Ilya Gruzberg, Andreas Ludwig, Christopher Mudry, Chetan Nayak, T. Senthil, Ben Simons, Martin Zirnbauer and especially Nick Read. I thank them for their patience in explaining most of this work to me. I would also like to thank Alexei Tsvelik for organizing such a nice conference. This work was supported by a DOE OJI Award, a Sloan Foundation Fellowship, and by NSF grant DMR-9802813.
|
warning/0006/cond-mat0006499.html
|
ar5iv
|
text
|
# Shot-noise suppression by Fermi and Coulomb correlations in ballistic conductors
## I Introduction
Nonequilibrium fluctuations of the electric current (shot noise) in mesoscopic conductors have received recently significant attention. In particular, the shot noise in scattering-free or ballistic conductors has been studied extensively both theoretically and experimentally, by focusing mainly on the suppression of noise by Fermi correlations in quantum point contacts under low temperatures, i.e., conductors with a small number of quantum modes.
On the other hand, when the ballistic transport is limited by a space charge, Coulomb correlations may also result in a shot-noise suppression. If the electron density injected into a ballistic conductor is low, the electron gas is nondegenerate, and Fermi statistical correlations are not efficient. For this case the Coulomb correlations are the main source of the shot-noise suppression, as has been demonstrated by Monte Carlo simulations, and subsequently analytically in a framework of the Vlasov system of equations. In nanoscale devices, however, the injected carriers are usually degenerate, which is due to a high level of contact doping and the elevated position of the Fermi level in the contact emitter. Therefore, it is of interest to consider the situation when both mechanisms, Fermi and Coulomb correlations, act togetherโthe case that is important not only from a fundamental, but also from an applied point of view and has attracted less attention so far. In Ref. the problem for a multimode degenerate conductor in the presence of a nearby gate has been posed, and the numerical results has been presented for a two-dimensional field-effect-transistor geometry. Monte Carlo simulations in a two-terminal geometry, which take into account the degenerate injection from the contacts and Coulomb correlations in the ballistic region, have been performed. The relative significance of each mechanism in the shot-noise suppression and the limiting values for the noise suppression factors of each mechanism still remain unclear, since the analytical theory has not been proposed.
It is the objective of the present paper to address the problem of a shot-noise suppression under the conditions of the interplay between Fermi and Coulomb correlations in two-terminal multimode ballistic conductors. To this purpose, we apply the recently developed analytical theory for space-charge-limited (SCL) ballistic conductors to the case of a Fermi-Dirac degenerate injection. Since we address the case of thick (in transversal dimensions) samples, the number of transversal modes (quantum channels) is large and the dimensionality of a momentum space of electrons is three-dimensional (3D), which makes a difference with the previous considerations of a one-channel or a few-channel quantum ballistic conductors (1D or quasi-1D momentum space). Our analysis goes beyond the linear-response regime and zero-temperature limit, โ the assumptions typically used to study a-few-channel conductors. In a semiclassical framework, for a multimode ballistic conductor, we have derived analytical formulas that determine the nonlinear $`I`$-$`V`$ characteristics, the current-noise spectral density, and the shot-noise suppression factors for each suppression mechanism in the limit of high biases. We show that the Fermi shot-noise-suppression factor is limited below by the value determined by the properties of the injecting contact (the ratio between the temperature and the Fermi energy), whereas the Coulomb noise suppression may be enhanced arbitrarily strong by extending the length of the ballistic sample with a simultaneous increase of bias (provided the transport remains ballistic). Therefore, the Coulomb suppression may be achieved much stronger even in samples with a high degree of an electron degeneracy.
The paper is organized as follows. In Sec. II we describe the semiconductor structure under consideration and discuss the main assumptions concerning the model. In Sec. III we introduce the electron distribution function over the longitudinal injection energy, found by integrating over the transversal modes. The analytical expression for the mean current is derived as a function of the self-consistent potential barrier height. Then, in the limit of high biases, the current-voltage characteristics beyond the Child approximation is obtained, which takes into account the degenerate Fermi-Dirac injection. In Sec. IV the analytical expression for the suppressed value of the shot-noise power is derived, in which the Fermi- and Coulomb-correlation contributions are distinguished. The results for a particular GaAs semiconductor SCL diode are presented in Sec. V. Finally, Sec. VI summarizes the main conclusions of the paper.
## II The physical model
We consider a two-terminal semiconductor ballistic sample with plane-parallel heavily doped contacts at $`x`$=0 and $`x`$=$`\mathrm{}`$. The structure may be considered as a $`n`$-$`i`$-$`n`$ SCL homodiode in which the current is determined by a charge injection from the contacts rather than by intrinsic carriers of the ballistic region. The applied bias $`U`$ between the contacts is assumed to be fixed by a low-impedance external circuit and does not fluctuate. In order to simplify the problem, we assume that due to the large difference in the carrier density between the contacts and the sample, and hence in the corresponding Debye screening lengths, all the band bending occurs in the ballistic base, and the relative position of the conduction band and the Fermi level does not change in the contacts. Therefore, when the bias is changed, the potential can vary exclusively inside the ballistic base, and the contacts are excluded from the consideration. The electron gas inside the contacts is assumed to be in thermal equilibrium. However, in contrast to the previous works, the Fermi level in respect to the bottom of the conduction band, denoted here $`\epsilon _F`$, may take not only negative, but positive values as well, i.e., the injected electrons may be either degenerate or nondegenerate, and follow, in general, the Fermi-Dirac distribution. Assuming the transversal size of the conductor sufficiently thick and high enough electron density, the electrostatic problem is considered in a one-dimensional plane geometry.
## III Distribution function and mean current
To describe the steady-state transport and low-frequency noise, we use a semiclassical Vlasov system of equations, which consists of the collisionless Boltzmann transport equation for the distribution function and the Poisson equation for the self-consistent electrostatic potential. Due to a stochastic nature of the injection, the distribution function and, consequently, the self-consistent potential both fluctuate in time. The nonuniform distribution of the injected carriers leads to the creation of the potential minimum $`\phi _m`$ at a position $`x=x_m`$. The potential minimum acts as a barrier for the electrons by reflecting a part of them back to the contact, thereby affecting the transport and noise properties. It is the potential minimum fluctuations that induce the long-range Coulomb interactions and lead to the suppression of the injected current fluctuations. We assume that the applied bias $`qU>5k_BT`$, where $`q`$ is the electron charge and $`T`$ is the temperature. From this follows that the current is determined by only one injecting contact (at $`x`$=0 for definiteness), and the electrons from this contact that are able to pass over the barrier and arrive at the receiving contact at $`x`$=$`\mathrm{}`$ are all absorbed with the probability 1, since the corresponding energy states are empty. All the electrons injected from the receiving contact are reflected back because of the high-bias condition. Their contribution to the current and noise is negligible. Another assumption on the bias is $`U_mU<U_{cr}`$, where $`U_m\phi _m`$ is the potential barrier height, and $`U_{cr}`$ is the bias at which the potential barrier vanishes. In this limit (โvirtual-cathode approximationโ), only the electrons that are able to pass over the fluctuating barrier contribute to the current and noise. The nonhomogeneous electron density along the ballistic region is determined by
$`N(x)={\displaystyle _{\mathrm{\Phi }_c}^{\mathrm{}}}F_c(\epsilon ){\displaystyle \frac{d\epsilon }{2\sqrt{\epsilon +\mathrm{\Phi }(x)\mathrm{\Phi }_c}}},`$ (1)
where $`\mathrm{\Phi }(x)=q\phi (x)q\phi _m`$ is the mean potential referenced to the minimum, with the value $`\mathrm{\Phi }_c\mathrm{\Phi }(0)`$ at the injecting contact. It is clear that in such a definition the contact potential is equal to the potential barrier height, $`\mathrm{\Phi }_c=qU_m`$. Note that $`F_c(\epsilon )`$ is the distribution function over the longitudinal kinetic energy $`\epsilon `$ at the injecting contact. Since during the ballistic motion only the longitudinal electron momentum may vary, the injection distribution function is averaged over the transversal momentum $`๐ค_{}`$:
$$F_c(\epsilon )=2\frac{\sqrt{2m}}{\mathrm{}}\frac{d๐ค_{}}{(2\pi )^d}f(\epsilon ,๐ค_{}),$$
(2)
where $`d`$ is the dimension of a momentum space, $`m`$ the electron effective mass, $`\mathrm{}`$ the Planck constant, $`f(\epsilon ,๐ค_{})`$ the occupation number of a quantum state, the factor 2 takes into account the spin variable, and the additional factor $`\sqrt{2m}/\mathrm{}`$ has been introduced for normalization convenience. Assuming that the number of transversal modes is large, the dimension of a momentum space $`d`$=$`3`$, and we can perform integration over the transversal states. Changing the variable of integration $`d๐ค_{}=(2\pi m/\mathrm{}^2)d\epsilon _{}`$, where $`\epsilon _{}`$ is the transverse electron energy, and taking into account that $`f(\epsilon ,\epsilon _{})=f_F(\epsilon +\epsilon _{})`$, with $`f_F(\epsilon )=\{1+\mathrm{exp}[(\epsilon \epsilon _F)/k_BT]\}^1`$ the Fermi-Dirac distribution, one gets
$$F_c(\epsilon )=\frac{N_c}{\sqrt{\pi k_BT}}\mathrm{ln}\{1+\mathrm{exp}[(\epsilon _F\epsilon )/k_BT]\},$$
(3)
where $`N_c=2(2\pi mk_BT)^{3/2}/(2\pi \mathrm{})^3`$ is the effective density of states. Integrating the distribution function (3) over the energy, one obtains the electron density injected from the contact,
$`N_0={\displaystyle _0^{\mathrm{}}}F_c(\epsilon ){\displaystyle \frac{d\epsilon }{2\sqrt{\epsilon }}}={\displaystyle \frac{N_c}{\sqrt{\pi }}}{\displaystyle _0^{\mathrm{}}}\mathrm{ln}(1+e^{\xi z^2})๐z,`$ (4)
where $`\xi \epsilon _F/k_BT`$ is the reduced Fermi energy. The injected electron density may also be expressed in a more familiar form
$`N_0={\displaystyle \frac{1}{2}}N_c_{1/2}(\xi ),`$ (5)
where $`_{1/2}`$ is the Fermi-Dirac integral of index $`1/2`$. Since the Fermi-Dirac integrals $`_j`$ of different indexes $`j`$ will be frequently used throughout the paper, their properties are summarized in the Appendix. Note that $`N_0`$ is half of the contact electron density $`N_c_{1/2}(\xi )`$, since only the electrons with positive momenta are injected into the sample.
Substituting the distribution (3) into Eq. (1) taken at $`x`$=$`x_m`$, one obtains the electron density at the potential minimum:
$`N_m={\displaystyle _{\mathrm{\Phi }_c}^{\mathrm{}}}F_c(\epsilon ){\displaystyle \frac{d\epsilon }{2\sqrt{\epsilon \mathrm{\Phi }_c}}}={\displaystyle \frac{1}{2}}N_c_{1/2}(\alpha ),`$ (6)
where $`\alpha =(\epsilon _F\mathrm{\Phi }_c)/k_BT`$ is the parameter characterizing the position of the Fermi energy with respect to the potential barrier. The density $`N_m`$ is an important parameter for computing the current noise, as will be seen below.
The steady-state current is obtained by
$$I=\frac{qA}{\sqrt{2m}}_{\mathrm{\Phi }_c}^{\mathrm{}}F_c(\epsilon )๐\epsilon ,$$
(7)
where $`A`$ is the cross-sectional area. Substituting the distribution function (3), one gets
$$I=I_F_0^{\mathrm{}}\mathrm{ln}(1+e^{\alpha y})๐y=I_F_1(\alpha ),$$
(8)
where $`I_F=4\pi qAm(k_BT)^2/(2\pi \mathrm{})^3`$ and $`_1`$ is the Fermi-Dirac integral (see the Appendix). It is seen that under the ballistic SCL conduction, the current is determined by the relative position of the Fermi energy and the potential barrier through the parameter $`\alpha `$. This is in contrast to the case of diffusive conductors, in which the current is determined by scattering strength. The parameter $`\alpha `$ summarizes the dependence of the current on the applied bias and the length of the conductor, since they both affect the potential barrier height, whereas the factor $`I_F`$ is independent of those characteristics.
Figure 1 illustrates the electric current as a function of $`\alpha `$ given by Eq. (8). When the Fermi energy is sufficiently below the potential barrier, $`\alpha <3`$, only the exponential tail of the contact distribution function is injected (nondegenerate injection limit). Under this condition, according to the approximate formulas for the Fermi-Dirac integrals (A5), the current becomes
$$II_Fe^\alpha .$$
(9)
When the Fermi energy is above the potential barrier by several $`k_BT`$, it is the degenerate injection limit and, by using Eq. (A6), one gets the approximate formula for the current
$$I\frac{1}{2}I_F\left(\alpha ^2+\frac{\pi ^2}{3}\right).$$
(10)
It is seen from Fig. 1, that formula (10) is accurate at $`\alpha >2`$. Note that the case of nondegenerate injection $`\alpha <3`$, may occur when the contact electron density is either nondegenerate or degenerate, depending on the position of the Fermi energy with respect to the conduction band edge characterized by the parameter $`\xi `$. For $`\xi <3`$, the contact electron density is nondegenerate, and this is the case of the Maxwell-Boltzmann injection, analyzed in detail in Ref. . Let us demonstrate that our formulas are in agreement with that case. Equation (5) gives the injected electron density $`N_0=\frac{1}{2}N_ce^\xi `$, and the current (9) is expressed as
$$I=I_{MB}e^{U_m/k_BT},$$
(11)
where $`I_{MB}=I_F(2N_0/N_c)=qAN_0\sqrt{2k_BT/\pi m}`$ is the emission current for the Maxwell-Boltzmann distribution \[compare with Eq. (46) of Ref. \]. For $`\xi >3`$ and $`\alpha <3`$, the injected electrons that pass over the barrier are nondegenerate, but the contact electrons are degenerate; hence the approximate formula (11) for the current is no longer valid, and one has to use a more general relation, Eq. (9).
It should be also noted that in the general case of a Fermi-Dirac injection, the contact emission current is $`I_0=I_F_1(\xi )`$. This is the maximum (saturation) current that is achieved when the applied bias $`UU_{cr}`$, the barrier vanishes ($`\mathrm{\Phi }_c=0`$, $`\alpha =\xi `$), and the conduction is no longer space-charge limited. The current in units of its saturation value is simply
$$\frac{I}{I_0}=\frac{_1(\alpha )}{_1(\xi )}.$$
(12)
It was shown in a previous paper that the asymptotic behavior of the current in SCL ballistic conductors obeys the Child law in the leading-order terms independently of the injection distribution:
$$I_{\mathrm{Child}}=\frac{4}{9}\kappa A\sqrt{\frac{2q}{m}}\frac{U^{3/2}}{\mathrm{}^2},$$
(13)
where $`\kappa `$ is the dielectric permittivity, and $`\mathrm{}`$ is the length of the ballistic conductor. However, this formula is only accurate at very high biases, in the range where the SCL conduction is difficult to maintain. This is a consequence of a rough approximation, in which the velocity spread of electrons at the potential minimum is neglected. To obtain a satisfactory good approximation at lower biases, it is necessary to keep the next-order terms that are specific with respect to the injection distribution. The general formula for an arbitrary injection function has been recently derived:
$$I=I_{\mathrm{Child}}\left(1+\frac{3}{\sqrt{qU}}\frac{_0^{\mathrm{}}F_c(ฯต+\mathrm{\Phi }_c)ฯต^{1/2}๐ฯต}{_0^{\mathrm{}}F_c(ฯต+\mathrm{\Phi }_c)๐ฯต}\right).$$
(14)
In our case of the injection distribution function (3), one finds the following expression
$$I=I_{\mathrm{Child}}\left(1+\frac{3\sqrt{\pi }}{2}\sqrt{\frac{k_BT}{qU}}\frac{_{3/2}(\alpha )}{_1(\alpha )}\right).$$
(15)
In the nondegenerate limit, $`\alpha <3`$, one obtains ($`_{3/2}/_1)1`$, and Eq. (15) leads to the Langmuir formula for the Maxwell-Boltzmann injection. In the opposite limit of high degeneracy, $`\alpha 1`$, one gets
$$II_{\mathrm{Child}}\left(1+\frac{8}{5}\sqrt{\frac{\epsilon _FqU_m}{qU}}\right),$$
(16)
which can be used to estimate the current for Fermi ballistic conductors beyond the Child approximation. Here, we remark that in the degenerate limit and at $`\epsilon _FqU_m`$, the current (16) is independent of temperature $`T`$ in both terms. For an arbitrary degree of degeneracy, the general expression (15) can be used.
## IV Current noise
To calculate the current noise, one has to define the partial injection current $`I_c(\epsilon )`$ at the contact and its fluctuation properties. From Eq. (8) it follows that
$$I_c(\epsilon )=\frac{I_F}{k_BT}\mathrm{ln}[1+e^{(\epsilon _F\epsilon )/k_BT}],$$
(17)
which corresponds to the current carried by electrons with injection (longitudinal) energies between $`\epsilon `$ and $`\epsilon +d\epsilon `$, giving after the integration the total emission current $`I_0=_0^{\mathrm{}}I_c(\epsilon )๐\epsilon `$.
The correlation function for the fluctuations of the partial injection currents may be written generally as
$$\delta I_c(\epsilon )\delta I_c(\epsilon ^{})=K(\epsilon )(\mathrm{\Delta }f)\delta (\epsilon \epsilon ^{}),$$
(18)
where $`\mathrm{\Delta }f`$ is the frequency bandwidth (we assume the low-frequency limit). For the particular case of Fermi 3D injection, the function $`K(\epsilon )`$ is determined by
$$K(\epsilon )=2q\frac{2qA}{\mathrm{}}\frac{d๐ค_{}}{(2\pi )^d}f(\epsilon ,๐ค_{})[1f(\epsilon ,๐ค_{})].$$
(19)
The integration over the transversal states may be performed explicitly by taking into account that $`f_F(1f_F)=k_BT(f_F/\epsilon )`$ and $`_0^{\mathrm{}}๐\epsilon _{}f_F(\epsilon +\epsilon _{})[1f_F(\epsilon +\epsilon _{})]=k_BTf_F(\epsilon )`$. This gives a simple expression:
$$K(\epsilon )=\frac{2qI_F}{k_BT}f_F(\epsilon ),$$
(20)
We remind the reader that $`\epsilon `$ is the longitudinal energy component. The Fermi factor $`1f_F`$ has disappeared after the integration, but the Fermi noise-suppression effect is present in Eq. (20). Indeed, in the degenerate limit $`\epsilon _Fk_BT`$, the partial current (17) is a linear function of energy, $`I_c(\epsilon )(\epsilon _F\epsilon )`$, at $`\epsilon 0`$ \[see Fig. 2(a)\]. This occurs because of the increase of the number of transversal states as the longitudinal energy $`\epsilon `$ decreases. Despite the increasing of the number of states, the shot-noise power per unit energy represented by the function (20) is constant at $`\epsilon \epsilon _F`$ \[Fig. 2(a)\]. As a result, $`K(\epsilon )/2qI_c(\epsilon )1/\xi 1`$ at $`\epsilon 0`$, indicating the noise suppression effect. In contrast, for nondegenerate case, both functions $`\mathrm{exp}[(\epsilon _F\epsilon )/k_BT]`$, and $`K(\epsilon )/2qI_c(\epsilon )1`$ \[Fig. 2(b)\], which leads to the Poisson noise. Additionally, we note that, since $`I_FT^2`$, the injection noise vanishes, $`K(\epsilon )0`$, in the limit $`T0`$.
The current-noise spectral density for the electron flow, when Coulomb correlations are disregarded, is given by
$`S_I^{\mathrm{uncor}}={\displaystyle _{\mathrm{\Phi }_c}^{\mathrm{}}}K(\epsilon )๐\epsilon .`$ (21)
Here, the integration is performed over the energies above the barrier height $`\mathrm{\Phi }_c`$, since only the electrons transmitted over the barrier contribute to the current noise at high biases. Substitution of expression (20) yields
$`S_I^{\mathrm{uncor}}=2qI_F\mathrm{ln}(1+e^\alpha )=2qI_F_0(\alpha ).`$ (22)
From this result we find the shot-noise-suppression factor caused by Fermi correlations
$$\mathrm{\Gamma }_F=\frac{S_I^{\mathrm{uncor}}}{2qI}=\frac{_0(\alpha )}{_1(\alpha )}.$$
(23)
This function is plotted in Fig. 3. It is clear that in the nondegenerate limit, one gets $`(_0/_1)1`$, and obviously $`\mathrm{\Gamma }_F1`$. An important feature is that $`\mathrm{\Gamma }_F`$ is a decreasing function of $`\alpha `$. In the degenerate limit, $`\alpha 1`$, it approaches the asymptotic behavior:
$$\mathrm{\Gamma }_F\frac{2}{\alpha +(\pi ^2/3\alpha )}.$$
(24)
It is seen from Fig. 3 that formula (24) is accurate at $`\alpha >3`$. The limiting minimal value for $`\mathrm{\Gamma }_F`$ occurs when the barrier vanishes ($`\alpha =\xi `$),
$$\mathrm{\Gamma }_F^{\mathrm{min}}=\frac{2}{\xi }=\frac{2k_BT}{\epsilon _F}.$$
(25)
The numerical factor in Eq. (25) depends on the dimensionality of a momentum space. By taking different values of $`d`$ in Eqs. (2) and (19), one can get $`\mathrm{\Gamma }_F^{\mathrm{min}}`$=$`ck_BT/\epsilon _F`$ with $`c`$=2 ($`d`$=3), $`3/2`$ ($`d`$=2), and 1 ($`d`$=1). In all the cases, the shot-noise Fermi suppression is determined by the ratio between the temperature of the injected electrons $`T`$ and their Fermi energy $`\epsilon _F`$. For a fixed $`\epsilon _F`$, the suppression may be enhanced by decreasing the temperature $`T0`$, but it is independent of the bias, the ballistic length and the other parameters of the conductor.
The current noise, which takes into account both Fermi and Coulomb correlations, is determined by
$$S_I=_{\mathrm{\Phi }_c}^{\mathrm{}}\gamma ^2(\epsilon )K(\epsilon )๐\epsilon ,$$
(26)
where the energy-resolved shot-noise-suppression factor
$`\gamma (\epsilon )`$ $`=`$ $`{\displaystyle \frac{3}{\sqrt{qU}}}[\sqrt{\epsilon \mathrm{\Phi }_c}\upsilon ],`$ (27)
and the constant $`\upsilon `$ for an arbitrary injection distribution is given by
$$\upsilon =\frac{N_m}{F_c(\mathrm{\Phi }_c)}$$
(28)
By using Eqs. (3) and (6), we find for the Fermi 3D injection
$`\upsilon ={\displaystyle \frac{\sqrt{\pi k_BT}}{2}}{\displaystyle \frac{_{1/2}(\alpha )}{_0(\alpha )}}.`$ (29)
Thus, for the current-noise power (26), after using Eqs. (20), (27), and (29), we find
$$S_I=\beta \mathrm{\hspace{0.17em}2}qI\frac{k_BT}{qU}.$$
(30)
In this formula, the constant $`\beta `$ is determined only by $`\alpha `$:
$`\beta (\alpha )=9\left(1{\displaystyle \frac{\pi }{4}}{\displaystyle \frac{[_{1/2}(\alpha )]^2}{_0(\alpha )_1(\alpha )}}\right),`$ (31)
To distinguish the noise suppression caused by different mechanisms, one can define the shot-noise-suppression factor due to a pure Coulomb suppression
$`\mathrm{\Gamma }_C={\displaystyle \frac{S_I}{S_I^{\mathrm{uncor}}}}=\beta {\displaystyle \frac{k_BT}{qU}}{\displaystyle \frac{_1(\alpha )}{_0(\alpha )}},`$ (32)
whereas the overall shot-noise-suppression factor becomes
$$\mathrm{\Gamma }=\mathrm{\Gamma }_C\mathrm{\Gamma }_F=\frac{S_I}{2qI}=\beta \frac{k_BT}{qU}.$$
(33)
It is seen that the current noise may be suppressed by both the temperature $`T`$ and the bias $`U`$. This is in contrast to the pure Fermi suppression (25), which is sensitive to $`T`$, but independent of the bias. The dependence on $`U`$ comes from the Coulomb correlations and originates from the function $`\gamma (\epsilon )`$. The coefficient $`\beta `$ is a parameter that depends on the degree of degeneracy, as follows from Eq. (31). For the Fermi 3D injection, $`\beta `$ is a decreasing function of $`\alpha `$ ranging between two limiting values (see Fig. 3): $`\beta _{min}<\beta <\beta _{\mathrm{MB}}`$, where $`\beta _{\mathrm{MB}}=9(1\pi /4)1.9314`$ is a limiting value in the nondegenerate limit (Maxwell-Boltzmann injection), and $`\beta _{min}`$=1 is a limiting value in the degenerate limit. For high degeneracy, the approximate formula for $`\beta `$ may be obtained by using the expansions for the Fermi-Dirac integrals (A6), one gets
$`\beta 1+{\displaystyle \frac{2}{3}}{\displaystyle \frac{\pi ^2}{\alpha ^2}},\alpha 1.`$ (34)
Figure 3 demonstrates the validity of such an approximation. The asymptotic behavior of the Coulomb suppression factor $`\mathrm{\Gamma }_C`$ in a high degenerate limit is obtained as
$`\mathrm{\Gamma }_C{\displaystyle \frac{1}{2}}{\displaystyle \frac{\epsilon _F\mathrm{\Phi }_c}{qU}},\epsilon _F\mathrm{\Phi }_ck_BT,`$ (35)
which takes the minimal value at $`U=U_{cr}`$:
$`\mathrm{\Gamma }_C^{\mathrm{min}}{\displaystyle \frac{\epsilon _F}{2qU_{cr}}}.`$ (36)
It should be emphasized that the difference between the two noise-suppression mechanisms is fundamental: While $`\mathrm{\Gamma }_F`$ cannot be decreased further by varying the parameters of the conductor, since its minimal value is fixed by the contact properties \[by the parameter $`\xi `$ as follows from Eq. (25)\]. In contrast, the factor $`\mathrm{\Gamma }_C`$ may be decreased by increasing the ballistic length $`\mathrm{}`$ of the conductor, since for longer conductors the critical bias $`U_{cr}`$ under which the barrier disappears is higher, and $`\mathrm{\Gamma }_C`$ may drop deeper. As a consequence, $`\mathrm{\Gamma }`$ may also attain much lower values. It is important to highlight, that in both nondegenerate and degenerate limits, the total shot-noise suppression factor $`k_BT`$, and can therefore be reduced by decreasing the temperature.
## V Example
To illustrate the results, consider the GaAs $`n`$-$`i`$-$`n`$ ballistic diode of length $`\mathrm{}`$=0.5$`\mu `$m at $`T`$=4 K. For this temperature and $`m`$=0.067$`m_0`$, the effective density of states is $`N_c6.7\times 10^{14}`$ $`\mathrm{cm}^3`$. Assuming the contact doping $`1.6\times 10^{16}\mathrm{cm}^3`$, the reduced Fermi energy $`\xi 10`$, and the contact electrons are degenerate. For this set of parameters, the Debye screening length associated with the contact degenerate electron density is approximately $`L_D=\sqrt{\kappa k_BT/[q^2N_c_{1/2}(\xi )]}14`$ nm. Since $`L_D\mathrm{}`$, the space-charge effects and, therefore, the Coulomb shot-noise suppression are important in a wide range of biases.
Let us introduce the normalized biases: $`V=qU/k_BT`$ and $`V_m=qU_m/k_BT`$. The calculation of the steady-state potential profile for different biases $`V`$ shows that the potential barrier varies from $`V_m11.2`$ at $`V`$=10 to $`V_m`$=0 at $`V`$=$`V_{cr}705`$ ($`U_{cr}243`$ mV) (see Fig. 4). In this range, the charge-limited conduction is controlled by the barrier height $`V_m`$, and by increasing the bias, one can observe the crossover from nondegenerate ($`\alpha =\xi V_m<1`$) to degenerate ($`2<\alpha <10`$) injection. This crossover is illustrated in Fig. 4, where the shot-noise suppression factors $`\mathrm{\Gamma }`$, $`\mathrm{\Gamma }_F`$, and $`\mathrm{\Gamma }_C`$ are plotted as functions of bias. Indeed, the Fermi suppression factor $`\mathrm{\Gamma }_F`$ varies from 1 at low biases to $`2/\xi 0.2`$ at high biases, in agreement with formulas (23)โ(25). Moreover, the factor $`\mathrm{\Gamma }`$ lies between two asymptotic lines: $`\beta _{\mathrm{MB}}(k_BT/qU)`$ at low biases (nondegenerate limit) and $`k_BT/qU`$ at high biases (degenerate limit), in agreement with Eq. (33) and the variation of $`\beta `$ in Fig. 3. The Coulomb correlation factor $`\mathrm{\Gamma }_C`$ decreases with bias up to the lowest value $`0.0078`$ at $`U`$=$`U_{cr}`$. After that value it increases sharply to 1 due to the disappearance of the potential barrier. The sharp increase of $`\mathrm{\Gamma }_C`$ at $`U=U_{cr}`$ is discontinuous in this asymptotic theory, which neglects the high-order terms in the expansions. The exact calculations give a smoother behavior. Note that both mechanisms essentially suppress shot noise at large $`U`$, but $`\mathrm{\Gamma }_C`$ is always much lower than $`\mathrm{\Gamma }_F`$ under SCL conditions.
## VI Summary
In conclusion, we have derived the analytical formulas that describe the mean current and the shot-noise power in degenerate space-charge-limited ballistic conductors. In the framework of a semiclassical Vlasov system of equations, which takes into account the fluctuations of the potential profile self-consistently, we have obtained a deep shot-noise suppression of more than two orders of magnitude caused by two independent mechanisms: Fermi and Coulomb correlations. The derived formulas clearly distinguish the shot-noise suppression factors caused separately by Fermi correlations (23), Coulomb correlations (32), and by the joint action of both \[Eq. (33)\].
We show that the Fermi shot-noise-suppression factor is limited below by the ratio between the temperature and Fermi energy of the contact electrons. The Coulomb noise-suppression factor, however, may attain much lower values $`\epsilon _F/2qU`$, because of its dependence on the applied bias $`Uk_BT/q`$. The asymptotic behavior of the overall shot-noise-suppression factor in a high degenerate limit was found to be $`k_BT/qU`$, independently of the material parameters. Finally, for the degenerate Fermi-Dirac injection, the asymptotic formula for the mean current beyond the Child approximation is proposed.
###### Acknowledgements.
We are grateful to V. A. Kochelap for valuable discussions. This work has been partially supported by the Generalitat de Catalunya, Spain, and the NATO linkage Grant No. HTECH.LG 974610.
## A Fermi-Dirac functions and their approximations
The Fermi-Dirac functions are encountered, whenever one wants to describe the electronic transport in degenerate semiconductor or metallic systems, and they are defined as
$$_j(\alpha )=\frac{1}{\mathrm{\Gamma }(j+1)}_0^{\mathrm{}}\frac{y^jdy}{1+e^{y\alpha }},$$
(A1)
where $`\mathrm{\Gamma }(j)`$ is the gamma function of the index $`j`$. For the expressions of this paper, the $`\mathrm{\Gamma }`$ functions take the values $`\mathrm{\Gamma }(\frac{3}{2})=\sqrt{\pi }/2`$, $`\mathrm{\Gamma }(\frac{5}{2})=3\sqrt{\pi }/4`$, $`\mathrm{\Gamma }(1)=\mathrm{\Gamma }(2)=1`$.
For positive indexes $`j`$, the Fermi-Dirac integrals can also be rewritten \[obtained by integrating by parts (A1)\]:
$$_j(\alpha )=\frac{j}{\mathrm{\Gamma }(j+1)}_0^{\mathrm{}}y^{j1}\mathrm{ln}(1+e^{\alpha y})๐y,$$
(A2)
A simple relation between the integrals of different orders is
$$_j(\alpha )=d_{j+1}/d\alpha .$$
(A3)
Unfortunately, these integrals cannot be resolved analytically except for the trivial cases:
$$\begin{array}{ccc}_0(\alpha )=\mathrm{ln}(1+e^\alpha ),& & j=0\\ _1(\alpha )=(1+e^\alpha )^1,& & j=1.\end{array}$$
(A4)
However, for small and large $`\alpha `$, one may use the approximate formulas for the nondegenerate limit, $`\alpha <2`$,
$$_j(\alpha )=e^\alpha j,$$
(A5)
and for the degenerate limit, $`\alpha 1`$,
$$_j(\alpha )=\frac{\alpha ^{j+1}}{(j+1)\mathrm{\Gamma }(j+1)}\left[1+\frac{\pi ^2}{6}\frac{j(j+1)}{\alpha ^2}+O\left(\frac{1}{\alpha ^4}\right)\right].$$
(A6)
|
warning/0006/cond-mat0006328.html
|
ar5iv
|
text
|
# Transition from Abelian to non-Abelian FQHE states
## Abstract
We study the transition from the Abelian multi-component $`(3,3,1)`$ quantum Hall state to the non-Abelian one component Pfaffian state in bilayer two dimensional electron systems. We show that tunneling between layers can induce this transition. At the transition points part of the degrees of freedom that describe the $`(3,3,1)`$ state disappear from the spectrum, and the system is correctly described by the Pfaffian state, with quasi-particles that satisfy non-Abelian statistics. The mechanism described in this work provides for a physical Hamiltonian interpretation of the algebraic projection from the $`(3,3,1)`$ to the Pfaffian state that has been discussed in the literature.
Even denominator states in double layer two dimensional electron systems (2DES) have been observed experimentally and are theoretically quite well understood . The two 2DES are separated by a potential barrier that, if high and thick enough, will inhibit both Coulomb interactions and tunneling between layers. If the barrier is made thinner, Coulomb interactions will become important even if tunneling is still suppressed. The relevant parameter to measure this effect is the ratio $`d/l_0`$ where $`d`$ is the interlayer separation and $`l_0`$ is the magnetic length. In real samples neither Coulomb interactions nor tunneling can be completely neglected. Therefore a very rich phase diagram can be constructed with Coulomb interlayer interaction (or alternatively the distance $`d`$) on one axis and the tunneling amplitude on the other.
We will concentrate here on systems in which a quantized Hall plateau exists at total filling fraction $`\nu =1/2`$. The phase diagram for these systems was first discussed by Halperin. He assumed that the actual spin of the electrons was polarized in the direction of the external field, and that the two layers were completely equivalent. A possible experimental realization for this system is a single wide quantum well in which the self consistent Coulomb potential creates a barrier in the middle of the well with maxima in the electron density at the two edges. Halperin suggested that for an intermediate range of distances $`d`$ and vanishing tunneling, the so called $`(3,3,1)`$ state should be a stable phase for the system.
The $`(3,3,1)`$ state is a correlated bilayer state which is basically stabilized by Coulomb interactions. It was shown that if the layer separation is large enough the state collapses into decoupled layers due to the fact that interlayer Coulomb interactions become negligible. The variational wave function describing this state, proposed by Halperin in the context of spinful systems has the form
$$\mathrm{\Psi }_{331}=\underset{i<j}{}(z_iz_j)^3\underset{i<j}{}(w_iw_j)^3\underset{i,j}{}(z_iw_j)e^{\frac{1}{4}{\scriptscriptstyle (|z_i|^2+|w_i|^2)}},$$
(1)
where $`z_i`$ and $`w_i`$ are the coordinates of the electrons in each plane. The first two factors represent the correlations within each layer, and the last one corresponds to the intralayer correlations. The $`\nu =1/2`$ state was observed experimentally and it was checked numerically that its properties are indeed well described by the $`(3,3,1)`$ wave function (1) .
It was also conjectured that a transition to a Pfaffian state should occur, within the range of distances $`d`$ for which the $`(3,3,1)`$ state is stable at vanishing tunneling, when the tunneling amplitude is made large enough. The Pfaffian state, proposed by Moore and Read , is a candidate for a fractional quantum Hall state at $`\nu =1/2`$ in single layer systems, or in general for $`\nu =1/q`$ where $`q`$ is an even number (were we working with bosons with strong repulsive interactions, $`q`$ would be an odd integer ). Its variational wave function is given by
$$\mathrm{\Psi }_{Pf}=\mathrm{Pfaff}(\frac{1}{z_iz_j})\underset{i<j}{}(z_iz_j)^qe^{\frac{1}{4}{\scriptscriptstyle |z|^2}},$$
(2)
where the Pfaffian is defined for a $`2N\times 2N`$ antisymmetric matrix whose elements are $`M_{ij}`$ by
$$\mathrm{Pfaff}(M_{ij})=\frac{1}{2^NN!}\underset{\sigma \epsilon S_{2N}}{}\mathrm{sgn}(\sigma )\underset{k=1}{\overset{N}{}}M_{\sigma (2k1),\sigma (2k)}$$
(3)
or as the square root of the determinant of $`M`$. It was shown that this wave function arises from applying Wickโs theorem to real fermion fields, or as the real space BCS wave function for pairing of spinless fermions.
The $`\nu =1/2`$ states were extensively studied in experiments in a wide single quantum well sample, varying the well width and sheet density. It was concluded that the state observed was the $`(3,3,1)`$, i.e. the Pfaffian state did not show up within the range of tunneling amplitude and thickness scanned in the experiments. The authors argued nevertheless that it should still appear in the phase diagram for larger tunneling.
Our goal is to explore the above mentioned transition between the $`(3,3,1)`$ and the Pfaffian states. Therefore, we consider a system in which the interlayer separation $`d`$ is kept fixed, while the tunneling amplitude between layers can be changed arbitrarily. In other words, we will be looking at Halperinโs phase diagram for a given value of the interlayer separation, such that if the tunneling amplitude vanishes, the $`(3,3,1)`$ state is the stable phase of the system.
We start with the usual chiral boson approach for the edge theory of the $`(3,3,1)`$ state (see e.g. ), that was recently reviewed in with the inclusion of tunneling between layers. We further include a chemical potential term for the electrons. In this case the original theory, written in terms of two chiral bosons (a $`c=2`$ central charge Conformal Field Theory (CFT)), can be mapped into an effective theory with one chiral boson and two Majorana fermions. We then study the phase diagram as a function of electron tunneling $`\lambda `$ and chemical potential $`\mu `$.
As we have already mentioned, given that the spacing between layers is kept fixed at a value such that both pahses are stable, the Pfaffian state could describe a double layer sample in the limit in which the tunneling amplitude between the layers is large enough so as the two species of electrons of the $`(3,3,1)`$ state become indistinguishable. Since the edge theory for the Pfaffian state can be described by a $`c=3/2`$ CFT , the question is then how does this process occur physically, i.e. how does the $`(3,3,1)`$ CFT with $`c=2`$ evolve to the Pfaffian CFT with $`c=3/2`$. Related to this, it has been shown that the $`(3,3,1)`$ edge theory can be seen as the enveloping theory for the non Abelian Pfaffian state. Indeed, there is an algebraic procedure by which the two elementary quasi-holes of the $`(3,3,1)`$ state merge into one in the Pfaffian state by getting rid of an Ising CFT factor from the original edge theory. However, an explicit mechanism implementing physically this procedure is, to our knowledge, still lacking. In this letter we address this issue and show that electron tunneling between layers is capable of implementing this projection. More precisely, when the tunneling amplitude and/or the chemical potential increase, there is a critical line in the $`(\lambda ,\mu )`$ plane at which one of the degrees of freedom that describes the original theory disappears. We furthermore show that the remaining degrees of freedom acquire the quantum numbers of the elementary excitations for the Pfaffian state and non-Abelian statistics emerges.
The edge theory for the $`(3,3,1)`$ state is described by the Hamiltonian
$$H=\frac{1}{4\pi }๐xV_{ij}:_xu_i_xu_j:,$$
(4)
where colons denote standard normal ordering. Here $`x`$ is the coordinate along the edge, the $`u_i`$ are chiral bosonic fields whose compactification radius is $`1`$, and $`V_{ij}`$ is a symmetric matrix whose coefficients depend on the confining potential and the interparticle interactions at the edge,
$$V=\left(\begin{array}{cc}v& g\\ g& v\end{array}\right).$$
(5)
The commutation relations for the bosonic fields are
$$[u_i(x,t),u_j(x^{},t)]=i\pi K_{ij}\mathrm{sgn}(xx^{}),$$
(6)
where K is a symmetric matrix which characterizes the topological properties of the system
$$K=\left(\begin{array}{cc}3& 1\\ 1& 3\end{array}\right).$$
(7)
There exists an orthogonal transformation that diagonalizes $`V`$ and $`K`$ simultaneously, after which the Hamiltonian eq. (4) simply reads
$$H=\frac{1}{4\pi }dx[v_c:(_x\varphi _c)^2:+v_n:(_x\varphi _n)^2:],$$
(8)
where $`v_c=4(v+g)`$ and $`v_n=2(vg)`$. Notice that the condition $`detV>0`$ must hold in order that both modes have the same chirality. $`\varphi _c`$ and $`\varphi _n`$ refer to charged and neutral modes respectively, which are chiral bosons with standard commutation relations
$$[\varphi _i(x,t),\varphi _j(x^{},t)]=i\pi \delta _{ij}\mathrm{sgn}(xx^{}).$$
(9)
The electron operators can be written in this basis as follows
$`\psi _{e1}`$ $``$ $`:e^{i(\sqrt{2}\varphi _c+\varphi _n)}:`$ (10)
$`\psi _{e2}`$ $``$ $`:e^{i(\sqrt{2}\varphi _c\varphi _n)}:`$ (11)
while the quasi-particle operators are
$`\psi _{qp1}`$ $``$ $`:e^{i(\frac{1}{\sqrt{8}}\varphi _c+\frac{1}{2}\varphi _n)}:`$ (12)
$`\psi _{qp2}`$ $``$ $`:e^{i(\frac{1}{\sqrt{8}}\varphi _c\frac{1}{2}\varphi _n)}:.`$ (13)
In reference the authors considered the problem of adding uniform electron tunneling to the edge theory. Here we will study the same problem adding also a chemical potential for the electrons. Therefore we add to the Hamiltonian the following perturbation terms
$`H^{}=`$ $`\mu _0{\displaystyle }dx[:\psi _{e1}^{}\psi _{e1}+\psi _{e2}^{}\psi _{e2}:]`$ (14)
$`+`$ $`\lambda _0{\displaystyle }dx[:\psi _{e1}^{}\psi _{e2}+\psi _{e2}^{}\psi _{e1}:].`$ (15)
Using the bosonic representation for the electron operators we can write
$`:\psi _{e1}^{}\psi _{e1}+\psi _{e2}^{}\psi _{e2}:`$ $``$ $`(i2\sqrt{2}a_0_x\varphi _ca_0^2:(_x\varphi _n)^2:)`$ (16)
$`:\psi _{e1}^{}\psi _{e2}+\psi _{e2}^{}\psi _{e1}:`$ $``$ $`:e^{i2\varphi _n}+e^{i2\varphi _n}:,`$ (17)
where $`a_0`$ is the UV cut-off. In terms of these bosons the total Hamiltonian can be decoupled into charged ($`H_c`$) and neutral ($`H_n`$) sectors given by
$`H_c`$ $`=`$ $`{\displaystyle }dx[{\displaystyle \frac{1}{4\pi }}v_c:(_x\varphi _c)^2:\mu _c:_x\varphi _c:]`$ (18)
$`H_n`$ $`=`$ $`{\displaystyle ๐x\frac{1}{4\pi }(v_n+\mu _n)}:(_x\varphi _n)^2:`$ (19)
$``$ $`{\displaystyle ๐x\lambda }:(e^{i2\varphi _n}+e^{i2\varphi _n}):,`$ (20)
where $`\mu _c,\mu _n\mu _0`$ and $`\lambda \lambda _0`$.
The properties of the charged sector are not changed by the perturbation since the new term is linear in derivatives and can be absorbed by a shift in the bare Hamiltonian.
As for the neutral mode, it proves useful to decompose it (through conformal embedding) in terms of two chiral Majorana fermions
$$:e^{i\varphi _n}:(\chi _1+i\chi _2).$$
(22)
The Hamiltonian then reads
$$H_n=\frac{1}{2}dx(i(v_n+\mu _n+\lambda _{eff}):\chi _1_x\chi _1:+i(v_n+\mu _n\lambda _{eff}):\chi _2_x\chi _2:),$$
(23)
where $`\lambda _{eff}\lambda `$ .
We see that the two chiral Majorana fermions behave as free fields, but acquire different velocities which are determined by the bare velocity of the neutral boson $`v_n`$, the tunneling amplitude $`\lambda _{eff}`$ and the chemical potential $`\mu _n`$. Moreover, each Majorana sector describes a (chiral) Ising CFT.
It is clear now that, assuming that the perturbative treatment of the interaction Hamiltonian (15) is valid, there are two lines in the $`(\lambda _{eff},\mu _n)`$ plane, given by $`\mu _n=(v_n\pm \lambda _{eff})`$, on which one of the Majorana velocities vanishes. Though this observation is immediate from eq. (23), the study of the emerging state is non-trivial and constitutes the main result in the present work. The key observation is that when one of these two conditions is satisfied, the corresponding Ising sector disappears from the spectrum and, as we shall see, the remaining degrees of freedom describe the physics of the Pfaffian state. In fact, the Hamiltonian density for the zero-velocity mode vanishes, therefore its energy-momentum tensor and hence its central charge vanish. In this way, the central charge of the original system (the $`(3,3,1)`$ state) decreases by $`1/2`$. The remaining system is described by one chiral boson and one Majorana fermion with total central charge $`c_{eff}=3/2`$, which is the correct value for describing the Pfaffian state.
To make sure that the projection procedure drives the system to the Pfaffian state, we now show how the electron and quasi-particle operators (11), (13) in the $`(3,3,1)`$ phase come to describe the corresponding operators in this new phase. To this end, we rewrite the original electron and quasi-particle operators in terms of the charged boson and the Ising primary fields (the Majorana fermions $`\chi _a`$, the spin (order) operators $`\sigma _a`$ and their duals (disorder) $`\mu _a`$, where $`a=1,2`$ labels the two Ising sectors). Therefore the electron operators for the $`(3,3,1)`$ phase in eq. (11) can be written as
$`\psi _{e1}:e^{i\sqrt{2}\varphi _c}(\chi _1+i\chi _2):`$ (24)
$`\psi _{e2}:e^{i\sqrt{2}\varphi _c}(\chi _1i\chi _2):.`$ (25)
The neutral components of the quasi-particle operators can be combined and represented in terms of the order and disorder fields $`\sigma _a`$ and $`\mu _a`$ as
$`:e^{i\varphi _n/2}+e^{i\varphi _n/2}:\sigma _1\sigma _2,`$ (26)
$`:e^{i\varphi _n/2}e^{i\varphi _n/2}:\mu _1\mu _2.`$ (27)
This identification has been proven in reference by a careful analysis of operator product expansions on both sides. Then the quasi-particle operators can be written as
$`\psi _{qp1}+\psi _{qp2}:e^{i\frac{1}{\sqrt{8}}\varphi _c}\sigma _1\sigma _2:,`$ (28)
$`\psi _{qp1}\psi _{qp2}:e^{i\frac{1}{\sqrt{8}}\varphi _c}\mu _1\mu _2:.`$ (29)
The vanishing of the energy-momentum tensor for one of the two Ising sectors at the critical line implements a coset construction. The essence of the coset is to project a sector out from the physical Hilbert space. In the case at hand the projected subspace corresponds to one of the Ising sectors (say $`a=2`$) of the $`(3,3,1)`$ theory .
It should be stressed at this point that the coset projection appears in a natural way within this context. Previous treatments advocating the coset mechanism for projecting out an Ising sector were performed without any connection to a Hamiltonian description.
An important question that remains to be answered is how this projection acts on the quasi-particle and electron operators. This projection can be seen as if all primaries in the projected sector become trivial (they have vanishing conformal weights). More precisely, the quasi-particle operators (13) degenerate into a single quasi-particle operator
$$\begin{array}{c}\psi _{qp1}:e^{i\frac{1}{\sqrt{8}}\varphi _c}(\sigma _1\sigma _2+\mu _1\mu _2):\hfill \\ \psi _{qp2}:e^{i\frac{1}{\sqrt{8}}\varphi _c}(\sigma _1\sigma _2\mu _1\mu _2):\hfill \end{array}\}\psi _{qp}^{\mathrm{Pfaff}}:e^{i\frac{1}{\sqrt{8}}\varphi _c}\sigma _1:$$
(30)
(there are indeed two possible dual descriptions in terms of $`\sigma `$ or its dual $`\mu `$) describing quasi-particle excitations over the Pfaffian ground state. They have charge $`e/4`$ and, more importantly, exhibit non-Abelian statistics.
Besides, the two original electron operators are projected onto
$$\begin{array}{c}\psi _{e1}:e^{i\sqrt{2}\varphi _c}:(\chi _11_2+i1_1\chi _2)\hfill \\ \psi _{e2}:e^{i\sqrt{2}\varphi _c}:(\chi _11_2i1_1\chi _2)\hfill \end{array}\}\psi _e^{\mathrm{Pfaff}}\pm :e^{i\sqrt{2}\varphi _c}::e^{i\sqrt{2}\varphi _c}:\chi _1\pm :e^{i\sqrt{2}\varphi _c}:,$$
(31)
that is the Pfaffian electron operator plus a four quasi-particle bound state $`:e^{i\sqrt{2}\varphi _c}:`$ (c.f. eq. (30)).
Once the electron and quasi-particle operators at the edge are known, the (bulk) wave functions for both the ground state and excited states can be constructed following , by computing suitable correlation functions of those operators. In this way one recovers the expression in eq. (2) for the Pfaffian ground state. The computation of the wave function for four quasi-holes over the ground state shows that the quasi-particle statistics is non-Abelian.
It is worth mentioning that non-Abelian statistics arises in this context due to the fact that one of the order-disorder fields becomes trivial. The corresponding computation with the full (non projected) quasi-particle operators gives the correct Abelian statistics in the $`(3,3,1)`$ phase.
In summary, we have shown that if we add to the edge theory for the $`(3,3,1)`$ state tunneling and chemical potential terms, there exists a critical line where part of the degrees of freedom that describe the system becomes unphysical and disappears from the spectrum. This is precisely the line where the characteristic properties of the $`(3,3,1)`$ state are lost. In turn, at these points of the parameter space, the electron and quasi-particle operators of the $`(3,3,1)`$ state can be mapped into the corresponding operators of the Pfaffian state, and the statistics of quasi-particles becomes non-Abelian. The mechanism described in this work provides for a physical interpretation of the algebraic projection from the $`(3,3,1)`$ to the Pfaffian state that has been discussed in the literature . The question that remains to be answered is whether the Pfaffian state corresponds to a stable phase, i.e. if the system remains in this state beyond the critical lines. An alternative treatment to the one presented here is eventually needed to resolve this issue within the framework of the edge theories. According to the phase diagram found by comparing the Pfaffian state bulk wave function with the real space BCS wave function for pairing of spinless fermions , the system should be in a Pfaffian phase somewhere beyond those lines.
Acknowledgements: We acknowledge useful discussions with P. Degiovanni, E. Fradkin, A. Honecker, A. Lugo, E.F. Moreno, P. Pujol and I. Todorov. This work is partially supported by CONICET, ANPCyT through grants No. 03-02249 and 03-03924 (AL), and Fundaciรณn Antorchas (Argentina) through grants No. A-13622/1-18, A-13622/1-106 and A-13740/1-64.
|
warning/0006/nucl-th0006074.html
|
ar5iv
|
text
|
# MKPH-T-00-08 Analysis of the low-energy ๐NN-dynamics within a three-body formalism *footnote **footnote *Supported by the Deutsche Forschungsgemeinschaft (SFB 443)
## I Introduction
The low-energy interaction of $`\eta `$-mesons with few-nucleon systems has recently become the subject of vigorous investigations. A great deal of attention has been devoted to the search for possible bound or resonance states in these systems. After an experimental study at Brookhaven , which casts some doubt upon the possibility of observing $`\eta `$-nuclei with relatively large mass numbers A$``$12, the attention was redirected to the interaction of $`\eta `$-mesons with light nuclei . Recent theoretical investigations have mainly been stimulated by the precise measurements of the photoproduction and hadronic reactions with $`\eta `$-mesons in final states. The obtained results are often interpreted as strong experimental hints that correlated states for light $`\eta `$-nuclear systems might exist (see e.g. ). The case in point is the strong energy dependence of the experimental cross sections observed near the $`\eta `$-production threshold .
In the present work we deal primarily with the dynamical properties of the $`\eta NN`$-system. The understanding of these properties is important for the following reasons. Firstly, the $`\eta NN`$-system is interesting in itself. It is the simplest $`\eta `$-nuclear system which admits an exact solution within the three-body formalism. Furthermore, it may be regarded as an example of a three-body system in which the pairwise driving forces are attractive and which may in principle develop a bound (or virtual) state, or a low-energy three-particle resonance near zero energy. Secondly, the study of $`\eta NN`$-scattering may serve as a promising tool for the investigation of the corresponding processes in complex nuclei. The possible existence of such correlated objects would point to the crucial significance of two-nucleon mechanisms in the $`\eta `$-interaction with nuclei. This fact in turn would require the revision of the simple first order approximation to the $`\eta `$-nuclear optical potential adopted by the currently available models .
The main features of the $`\eta NN`$-dynamics in a quasideuteron state of spin, parity and isospin $`(J^\pi ;T)=(1^{};0)`$ are already elucidated in the literature. Ueda was the first to carry out an extensive calculation of $`\eta d`$-scattering within a three-body approach and found that the $`\eta d`$-system can form a quasi-bound state with a mass of 2430 MeV and a rather small width of about 10-20 MeV. Other calculations, also looking for bound or resonant $`\eta d`$-states within the Faddeev theory, have recently been reported by Shevchenko et al. as well as by Garcilazo and Peรฑa . The authors of Ref. confirmed qualitatively the results of and fixed the values of the $`\eta N`$-scattering length for which the previously bound $`\eta d`$-state becomes a low-lying $`s`$-wave resonance. In Ref. the possible existence of $`\eta d`$ bound states within the different $`\eta N`$\- and $`NN`$-interaction models is also studied. As for the experimental investigations, we would like to mention the measurements presented in for the reaction $`np\eta d`$ where a visible increase of the $`\eta `$-meson yield over the mere phase-space calculation was observed near threshold. Recently, Metag et al. have reported new results for the cross section $`\gamma d\eta X`$. These authors also note a strong enhancement of events in the energy region of a few MeV above threshold. Such a feature cannot be explained within the truncated multiple scattering approach developed in , where only single $`\eta N`$\- and $`NN`$-rescatterings were included in the calculation as the most important corrections to the impulse approximation. This discrepancy is not surprising, since the validity of the latter approach is associated with the short range nature of the $`\eta N`$\- and $`NN`$-interactions in relation to the characteristic internucleon distance in the deuteron. In this situation, when thinking about the whole picture of the $`\eta NN`$ dynamics, one may assume that, when the attractive forces pull all particles together so that the two-body potentials overlap, qualitatively new features of the resulting $`\eta NN`$-interaction may be expected.
In this paper we will study the question whether the dynamical properties of the $`\eta NN`$-system allow the existence of a bound (or virtual) state or a three-body resonances in the low-energy region. Although this question has already been covered partially in the above mentioned works we would like to reinvestigate it using a fundamentally different method, based on a search for poles of the scattering matrix in the complex energy plane. It possesses several important advantages over the conventional procedure of solving the on-shell Faddeev equation. Besides being more convenient for determining the exact position of $`S`$-matrix poles, the method provides additional insights into the source of appearance of these poles close to the physical domain. Furthermore, this approach allows us to investigate the $`(J^\pi ;T)=(0^{};1)`$ state which is not touched upon in the literature to the best of our knowledge. As was established in , this configuration plays the dominant role in the final state of the $`\gamma d\eta X`$ reaction near threshold. Therefore, if one attributes the anomalous behaviour of the $`\gamma d\eta X`$ cross section reported in to a near lying $`S`$-matrix pole, one has to search it in the $`(J^\pi ;T)=(0^{};1)`$ state. Broadly speaking, in view of the exclusive character of the measurements , the marked enhancement of the near threshold yield can be assigned to the $`\eta d`$ final state. But the simple estimation shows that in order to explain the observed results one would need an enhancement factor of about 30 what in our opinion seems unlikely. In this context, large attention is focussed on the $`(J^\pi ;T)=(0^{};1)`$ configuration in this paper.
Our principal tool is the three-body approach realized within the Alt-Grassberger-Sandhas formalism . Since the search for the $`S`$-matrix poles requires the analytical continuation of the dynamical equation into the nonphysical sheets of the Riemann surface, the analytical form for the driving two-body interactions is needed. For this reason, we use a simple separable potential of rank one and restrict our consideration to only $`s`$-waves in the two-body subsystems. This approach is well justified by the empirically established $`s`$-wave dominance of the low-energy $`\eta N`$\- and $`NN`$\- scattering. There exist strong experimental and theoretical evidences that the low-energy $`\eta N`$-interaction is dominated by the formation of the $`S_{11}`$(1535) resonance. Analogously, the $`{}_{}{}^{1}S_{0}^{}`$ and $`{}_{}{}^{3}S_{1}^{}`$ poles determine to a large extent the nucleon-nucleon low-energy interaction. Thus the $`s`$-wave isobars in the $`\eta N`$ and $`NN`$ two-body channels are expected to be the main source of the $`\eta NN`$-forces. Therefore we hope that our separable model, though being quite simple, will reproduce the major features of the $`\eta NN`$-dynamics.
In Sect. II we briefly describe the three-body scattering formalism pertinent to the present problem. Some details connected with the driving two-body interactions as well as the main calculational formulas are given in Sect. III. In Sect. IV the procedure of analytical continuation of the scattering equation into nonphysical energy sheets is described and the strategy for the search of the $`S`$-matrix poles is presented. The discussion of our main results and the conclusions are presented in the last two sections.
## II General formalism
In this section we will briefly review the properties of the three-body equation which will concern us in the remainder of this paper. Our starting point is the conventional three-body scattering theory in AGS form . The three channels comprising two interacting particles and one spectator are labeled according to the number of the spectator. As already mentioned in the introduction, we take for each channel as the off-shell two-body $`t`$-matrix a rank-one ansatz
$$t_k(E)=|f_k\tau _k(E)f_k|,$$
(1)
and we will call such an interacting pair in the following โisobarโ. Here $`|f_k`$ is the vertex function, and the isobar propagator $`\tau _k`$ is defined as
$$\tau _k(E)=\frac{\gamma _k}{1\gamma _kf_k|G_k^{(2)}(E)|f_k},$$
(2)
where $`G_k^{(2)}(E)`$ denotes the free two-body Greenโs function with $`E`$ as total c.m. energy of the two-particle system. The separable form (1) enables one to reduce the three-body problem to a set of coupled effective two-body equations for the transition amplitudes $`X_{ij}`$
$$X_{ij}(W)=(1\delta _{ij})Z_{ij}(W)+\underset{k=1}{\overset{3}{}}(1\delta _{ik})Z_{ik}(W)\tau _k(E_k)X_{kj}(W),(i,j=1,2,3),$$
(3)
where $`W`$ denotes the total three-body energy while the isobar propagator $`\tau _k`$ depends explicitly on the invariant energy $`E_k`$ of the two-body subsystem, which is a function of $`W`$ and the kinetic energy of the spectator. The energy dependent driving terms are
$$Z_{ij}(W)=f_i|G(W)|f_j,$$
(4)
with $`G(W)`$ being the free three-body Greenโs function. When evaluating equation (4) in momentum space and making a partial wave decomposition, it becomes a set of one-dimensional integral equations of Fredholm type. It is known that for an inhomogeneous Fredholm equation to be singular, in which case the transition matrix $`X_{ij}`$ has a pole, it is necessary and sufficient for the corresponding homogeneous equation
$$F_{ij}(W)=\underset{k=1}{\overset{3}{}}(1\delta _{ik})Z_{ik}(W)\tau _k(E_k)F_{kj}(W),(i,j=1,2,3)$$
(5)
to have a nonzero solution for the same value of the parameter $`W`$. Thus the problem of searching for poles of the scattering matrix reduces to that of finding those values $`W`$ for which the Fredholm determinant $`D(W)`$ of (3) vanishes
$$D(W)=0.$$
(6)
We now turn to the $`\eta NN`$-system for which we adopt the labeling $`k=1,2`$ for a nucleon spectator and $`k=3`$ for the $`\eta `$ spectator, and furthermore change the channel notation as
$$\begin{array}{cc}1=2N^{}& \text{ for the }\eta N\text{-isobar plus spectator nucleon},\hfill \\ 3d& \text{ for the }NN\text{-isobar plus spectator meson}.\hfill \end{array}$$
(7)
The identity of the channels 1 and 2 reduces the 3$`\times `$3 system (5) to a 2$`\times `$2 system which has the following operator form in our new notation
$`F_{dd}=2Z_{dN^{}}\tau _N^{}F_{N^{}d},`$ (8)
$`F_{N^{}d}=Z_{N^{}d}\tau _dF_{dd}+Z_{N^{}N^{}}\tau _N^{}F_{N^{}d},`$ (9)
$`F_{dN^{}}=2Z_{dN^{}}\tau _N^{}F_{N^{}N^{}},`$ (10)
$`F_{N^{}N^{}}=Z_{N^{}d}\tau _dF_{dN^{}}+Z_{N^{}N^{}}\tau _N^{}F_{N^{}N^{}}.`$ (11)
A close examination of (8) through (11) reveals that we have two decoupled sets of coupled equations sharing the identical integration kernel. Indeed, the coupled equations (8) and (9) are transformed into (10) and (11) by the substitutions $`F_{dd}F_{dN^{}}`$ and $`F_{N^{}d}F_{N^{}N^{}}`$. Therefore, it is sufficient to consider only one set for which we choose the second one. After inserting (10) into (11) the former equation may be written in the closed form
$$F_{N^{}N^{}}=(Z_{N^{}N^{}}+2Z_{N^{}d}\tau _dZ_{dN^{}})\tau _N^{}F_{N^{}N^{}}.$$
(12)
It is diagrammatically presented in Fig. 1. In analogy to a homogeneous Lippmann-Schwinger equation, we see that the driving term $`Z_{N^{}N^{}}`$ plays the role of a meson exchange $`NN^{}`$-potential while the second term in brackets gives the mechanism associated with intermediate $`NN`$-interaction where the $`\eta `$ acts as a spectator.
## III Two-body ingredients
Now we will specify the separable $`\eta N`$ and $`NN`$ scattering matrices which determine the driving two-body forces in our model. Since we restrict the pair-interaction to $`s`$-waves only, the vertex functions have a simple structure determined by two parameters, the coupling constant $`g_k`$ and the cut-off $`\beta _k`$
$$\stackrel{}{p}|f_k=f_k(p)=g_kF_k(p)\text{ with }F_k(p)=\frac{\beta _k^2}{\beta _k^2+p^2},$$
(13)
where $`\stackrel{}{p}`$ denotes the relative momentum of the interacting pair.
For the $`s`$-wave $`t`$-matrix of the $`NN=d`$ isobar
$$t_d(p,p^{},E_d)=f_d(p)\tau _d(E_d)f_d(p^{}),$$
(14)
the following parametrization has been used
$$g_d^2=\frac{16\pi a}{a\beta _d2},\gamma _d=\frac{1}{2M_N},$$
(15)
where $`a`$ is the $`NN`$-scattering length. For the propagator $`\tau _d`$ one then obtains
$$\tau _d(E_d)=\frac{1}{2M_N}\left[1+\frac{g_d^2\beta _d^3}{16\pi \left(i\beta _d+\sqrt{M_N(E_d2M_N)}\right)^2}\right]^1.$$
(16)
The $`NN`$-interaction parameters were taken from the low-energy $`np`$-scattering fit of Yamaguchi
$$\beta _d=1.4488\text{fm}^1,a=\{\begin{array}{cc}\hfill 5.378\text{fm}& \text{for the }{}_{}{}^{3}S_{1}^{}\text{-state,}\hfill \\ \hfill 23.690\text{fm}& \text{for the }{}_{}{}^{1}S_{0}^{}\text{-state.}\hfill \end{array}$$
(17)
Analogously, we use for the $`\eta N`$=$`N^{}`$ $`t`$-matrix the following ansatz
$$t_N^{}(p,p^{},E_N^{})=f_N^{}^{(\eta )}(p)\tau _N^{}(E_N^{})f_N^{}^{(\eta )}(p^{}).$$
(18)
Here the $`N^{}`$-propagator is given in the form
$$\tau _N^{}(E_N^{})=\left(E_N^{}M_0\mathrm{\Sigma }_\pi (E_N^{})\mathrm{\Sigma }_\eta (E_N^{})\right)^1,$$
(19)
where $`M_0`$ denotes the bare mass of the $`S_{11}`$(1535)-resonance. The resonance self energies associated with the couplings to the $`\pi N`$ and $`\eta N`$ channels are determined by
$$\mathrm{\Sigma }_j(E_N^{})=\frac{1}{2\pi ^2}\underset{0}{\overset{\mathrm{}}{}}\frac{f_N^{}^{(j)2}(p)}{E_N^{}E_N(p)\omega _j(p)+iฯต}\frac{p^2dp}{2\omega _j(p)},$$
(20)
where $`\omega _j(p)=\sqrt{m_j^2+p^2}`$ denotes the energy of meson โ$`j`$โ and $`E_N(p)=M_N+p^2/2M_N`$ is the nonrelativistic total nucleon energy. The meson-$`N^{}`$ vertices are defined by $`(j=\pi ,\eta )`$
$$f_N^{}^{(j)}(p)=g_N^{}^{(j)}F_N^{}^{(j)}(p),\text{with}F_N^{}^{(j)}(p)=\frac{\beta _N^{}^{(j)2}}{\beta _N^{}^{(j)2}+p^2}.$$
(21)
The $`N^{}`$ parameters were chosen in such a way that the main ratios of the hadronic decays of the $`S_{11}`$(1535) resonance are reproduced. We have taken
$$g_N^{}^{(\eta )}=2.0,g_N^{}^{(\pi )}=1.5,\beta _N^{}^{(\eta )}=6.5\text{fm}^1,\beta _N^{}^{(\pi )}=4.5\text{fm}^1.$$
(22)
The bare $`S_{11}`$ mass $`M_0`$ was determined by the condition
$$M_0+\mathrm{}e\left(\mathrm{\Sigma }_\pi (M^{})+\mathrm{\Sigma }_\eta (M^{})\right)=M^{},$$
(23)
where $`M^{}=1535`$ MeV is the mass of the dressed isobar. The choice (22) gives for the total and partial decay widths at the resonance position
$$\mathrm{\Gamma }=150\text{MeV},\frac{\mathrm{\Gamma }_{\eta N}}{\mathrm{\Gamma }}=\frac{\mathrm{\Gamma }_{\pi N}}{\mathrm{\Gamma }}=0.5,$$
(24)
which is reasonably consistent with the values given in the 1998 Particle Data Group listings .
For the actual evaluation of (12) we use a natural representation in which the total angular momentum $`J`$ and the total isospin $`T`$ are diagonal. The $`\eta NN`$ wave function is expanded into a complete set of the following basis states
$`|\stackrel{}{q};JM_JTM_T`$ $`=`$ $`|\{(\tau _i\tau _j)t_k\tau _k\}TM_T{\displaystyle \underset{LM_L}{}}{\displaystyle \underset{SM_S}{}}C_{LM_LSM_S}^{JM_J}|\{(\sigma _i\sigma _j)s_k\sigma _k\}SM_S|q;LM_LY_{LM_L}(\widehat{q}).`$ (25)
Here $`\sigma _i`$ and $`\tau _i`$ denote the spin and isospin of the individual particles forming the isobar with the corresponding quantum numbers $`s_i`$ and $`t_i`$. The partial waves are normalized as
$$q;LM_L|q^{};L^{}M_L^{}=\frac{1}{q^2}\delta (qq^{})\delta _{L^{}L}\delta _{M_L^{}M_L}.$$
(26)
In (25) we have already taken into account that the two-particle forces act only in $`l=0`$ states. The momentum $`\stackrel{}{q}`$ characterizes the relative motion of the correlated $`(ij)`$-pair and a spectator particle. It is projected onto the partial waves specified by the total orbital momentum $`L`$ with projection $`M_L`$. Taking into account the parametrization adopted for the two-body vertices, the expansion (25) turns the operator equation (12) into a set of one-dimensional integral equations for the partial waves $`|q;LM_L`$. In the following we consider only states with total orbital angular momentum $`L=0`$. Then the two allowed configurations are $`(J^\pi ;T)=(0^{};1)`$ and $`(1^{};0)`$. Consequently, only the $`{}_{}{}^{1}S_{0}^{}`$ $`NN`$-state contributes to the three-body state $`(0^{};1)`$, and correspondingly the $`{}_{}{}^{3}S_{1}^{}`$ $`NN`$-state to $`(1^{};0)`$.
In order to apply the representation (25) to Eq. (12), we only need the spin-isospin recoupling coefficients between the states $`j`$ and $`k`$ for which one has
$`\{(\sigma _i\sigma _j)s_k\sigma _k\}SM_S|\{(\sigma _k\sigma _i)s_j\sigma _j\}S^{}M_S^{}`$ $`=`$ $`\delta _{S^{}S}\delta _{M_S^{}M_S}\sqrt{(2s_j+1)(2s_i+1)}W(\sigma _i\sigma _k\sigma _jS;s_js_k),`$ (27)
$`\{(\tau _i\tau _j)t_k\tau _k\}TM_T|\{(\tau _k\tau _i)t_j\tau _j\}T^{}M_T^{}`$ $`=`$ $`\delta _{T^{}T}\delta _{M_T^{}M_T}\sqrt{(2t_j+1)(2t_i+1)}W(\tau _i\tau _k\tau _jT;t_jt_k),`$ (28)
where $`W(j_1j_2j_3j_4;j_5j_6)`$ denotes the standard Racah coefficient.
Taking the actual quantum numbers of the participating particles, we obtain the following explicit form of the driving terms appearing in (12)
$$\stackrel{}{p};JT|Z_{N^{}N^{}}|\stackrel{}{p}^{};JT=V_\eta (p,p^{},W)+\chi V_\pi (p,p^{},W).$$
(29)
Here the spin-isospin coefficients are $`\chi =1/3`$ for $`(0^{};1)`$ and $`\chi =1`$ for $`(1^{};0)`$. The meson-exchange potential acting in the $`L=0`$ wave reads
$$V_j(p,p^{},W)=\frac{g_N^{}^{(j)2}}{8\pi }\underset{1}{\overset{+1}{}}\frac{F_N^{}^{(j)}\left(\stackrel{}{p}^{}+\frac{M_N}{M_N+m_j}\stackrel{}{p}\right)F_N^{}^{(j)}\left(\stackrel{}{p}+\frac{M_N}{M_N+m_j}\stackrel{}{p}^{}\right)}{2\omega _j(|\stackrel{}{p}+\stackrel{}{p}^{}|)\left(WE_N(p)E_N(p^{})\omega _j(|\stackrel{}{p}+\stackrel{}{p}^{}|)\right)}d(\widehat{p}\widehat{p}^{}),$$
(30)
with $`(j=\pi ,\eta )`$. For simplicity we use the nonrelativistic relative meson-nucleon momenta in the arguments of the regularization vertex form factors. The driving term associated with the intermediate $`NN`$-interaction has the form, using $`E_d=W\omega _\eta (q)q^2/4M_N`$ as the total c.m. energy of the interacting $`NN`$-pair,
$`\stackrel{}{p};JT|Z_{N^{}d}\tau _dZ_{dN^{}}|\stackrel{}{p}^{};JT`$ $`=`$ $`V_d(p,p^{},W)`$ (31)
$`=`$ $`{\displaystyle \frac{2}{\pi }}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^2dq}{2\omega _\eta (q)}}V_N(p,q,W)\tau _d\left(W\omega _\eta (q)q^2/4M_N\right)V_N(p^{},q,W),`$ (32)
where the nucleon-exchange potential is given by
$$V_N(p,p^{},W)=\frac{g_N^{}^{(\eta )}g_d}{8\pi }\underset{1}{\overset{+1}{}}\frac{F_N^{}^{(\eta )}\left(\stackrel{}{p}^{}+\frac{m_\eta }{M_N+m_\eta }\stackrel{}{p}\right)F_d\left(\stackrel{}{p}+\frac{\stackrel{}{p}^{}}{2}\right)}{WE_N(p)\omega _\eta (p^{})E_N(|\stackrel{}{p}+\stackrel{}{p}^{}|)}d(\widehat{p}\widehat{p}^{}).$$
(33)
The integrals in (30) and (33) can be evaluated analytically. The corresponding expressions, being rather cumbersome, are listed in the appendix. Finally, we present the partial wave representation of our basic homogeneous equation (12)
$$F_{N^{}N^{}}(p,W)=\frac{2}{\pi }\underset{0}{\overset{\mathrm{}}{}}V_{N^{}N^{}}(p,p^{},W)\tau _N^{}(WE_N(p^{})p^{\mathrm{\hspace{0.17em}2}}/2M^{})F_{N^{}N^{}}(p^{},W)p^{}{}_{}{}^{2}dp^{},$$
(34)
where the notation
$$V_{N^{}N^{}}(p,p^{},W)=2V_d(p,p^{},W)+V_\eta (p,p^{},W)+\chi V_\pi (p,p^{},W)$$
(35)
is introduced, and the invariant mass of the $`N^{}`$-isobar in (35) was evaluated as $`E_N^{}=WE_N(p^{})p^2/2M^{}`$. Anticipating a later result, we note that the contribution from the pion exchange potential $`V_\pi `$ in (35) is practically insignificant, and almost the whole attraction in the $`\eta NN`$-system comes from the first two terms with approximately equal strengths.
The method for searching the zeros of the Fredholm determinant $`D(W)`$ is to approximate the integral in Eq. (34) by a finite sum, transforming it into an ordinary matrix equation. Then (6) may be written as an algebraic equation
$$det\left|\delta _{ij}\frac{2}{\pi }p_j^2C_jV_{N^{}N^{}}(p_i,p_j,W)\tau _N^{}\left(WE_N(p_j)\frac{p_j^2}{2M^{}}\right)\right|=0.$$
(36)
Here $`C_j`$ are the weights for the chosen quadrature (in the present calculation we have chosen the Gauss quadrature).
## IV Structure of the Riemann surface and continuation into nonphysical sheets
Our main concern is to find the zeros of the Fredholm determinant in the complex energy plane resulting in poles of the scattering matrix. The structure of the manifold Riemann surface for the $`\eta NN`$-system in $`(J^\pi ;T)=(1^{};0)`$ channel is presented in Fig. 2. Shown are the physical sheet $`\mathrm{\Pi }_1`$ and the nearest nonphysical sheets $`\mathrm{\Pi }_2`$, $`\mathrm{\Pi }_3`$ directly adjacent to $`\mathrm{\Pi }_1`$. The sheet $`\mathrm{\Pi }_1`$ has a conventional analytical structure. Namely, it may have poles corresponding to a possible formation of three-body bound states, and the unitarity right-hand cuts. The former stem from the Cauchy type integrals inherent in the integration kernel in (34). There are two main cuts which determine the structure of the Riemann surface for the three-body problem in the energy region of interest:
(i) A two-particle cut, beginning at the two-body threshold $`W_{\eta d}=2M_N+m_\eta +ฯต_d`$ where $`ฯต_d<0`$ is the deuteron binding energy. The cut arises from the propagator $`\tau _d(E_d)`$ which has a pole when $`E_d2M_N=ฯต_d`$. This pole is associated with the elastic $`\eta `$-deuteron scattering and gives the well known two-body contribution to the common three-body unitarity relations. The corresponding square-root branch point $`W=W_{\eta d}`$ develops the two-sheet structure $`\{\mathrm{\Pi }_1,\mathrm{\Pi }_2\}`$ typical for conventional two-body scattering.
(ii) A three-body cut, starting at the three-body threshold $`W_{\eta NN}=2M_N+m_\eta `$. This cut is induced by the two-body right-hand cut in $`\tau _d`$ corresponding to the $`dNN`$ break-up and by the similar cut in $`\tau _N^{}`$ associated with the decay $`N^{}\eta N`$. The corresponding nonphysical sheet is denoted as $`\mathrm{\Pi }_3`$ in Fig. 2. Furthermore, the driving terms $`V_\pi `$, $`V_\eta `$, and $`V_N`$ in (30) through (33) contain the well known logarithmic singularities which are analogous to the dynamical (left-hand) singularities of the nucleon-nucleon OBE potential and correspond to the exchange of a real particle which becomes possible when $`WW_{\eta NN}`$. For some values of $`p`$ and $`p^{}`$ these singularities pinch the real axis producing the additional three-body cut beginning at the branch point $`W=W_{\eta NN}`$. In the partial wave representation, this point is of logarithmic type and gives rise to an infinite number of sheets at $`W=W_{\eta NN}`$. The structure of the Riemann surface associated with the logarithmic singularities is in itself of no physical interest and is therefore not presented in Fig. 2.
All sheets are connected as depicted in Fig. 2. The sheet $`\mathrm{\Pi }_2`$ may contain the poles which show-up as correlated two-body $`\eta d`$ states (virtual or resonant). The analogous states which may be observed in the three particle scattering processes are located on the sheet $`\mathrm{\Pi }_3`$. In the absence of an actual three-body scattering experiment, these states may occur as final states in reactions with, e.g., deuteron break-up, such as $`\gamma d\eta np`$. Following the terminology accepted in the literature we call $`\mathrm{\Pi }_2`$ and $`\mathrm{\Pi }_3`$ as two-body and three-body sheets, respectively.
The singularity structure outlined above arises from the free three-body propagators and the poles associated with the bound states of two-body subsystems. It is also presented in detail in Refs. in relation to the three-nucleon problem. The Riemann surface for the $`\eta NN`$ scattering matrix is formally more complicated due to the following reasons:
(i) Since there is a rather strong coupling between the $`\eta N`$ and $`\pi N`$ channels in the energy region of the $`S_{11}`$(1535) resonance, all sheets depicted in Fig. 2 have an additional cut beginning at $`\pi NN`$ thresholdDue to the spin-isospin selection rules the $`\pi d`$ channel does not appear in the configurations $`(J^\pi ;T)=(1^{};\mathrm{\hspace{0.17em}0})`$ and $`(0^{};\mathrm{\hspace{0.17em}1})`$ considered in this paper. Therefore no additional two-body cut starting at the $`\pi d`$ threshold is present.. The typical structure of the Riemann surface associated with the $`\eta N\pi N`$ coupling is presented in Fig. 3.
(ii) The three-body sheet $`\mathrm{\Pi }_3`$ has an additional square-root branch point associated with the quasi-two-body $`NN^{}`$-threshold. Treating the bare $`N^{}`$ mass as noted in Sect. IV (see Eq. (23)), we keep the corresponding complex cut well away from the relevant energy region and, therefore, will ignore it in the following considerations.
The structure of the Riemann sheet for the $`\eta NN`$ scattering matrix in the $`(J^\pi ;T)=(0^{};1)`$ state is different from that for $`(J^\pi ;T)=(1^{};\mathrm{\hspace{0.17em}0})`$. Namely, due to the virtual character of the $`{}_{}{}^{1}S_{0}^{}`$ pole, the corresponding two-body sheet $`\mathrm{\Pi }_2`$ is โgluedโ not directly to $`\mathrm{\Pi }_1`$ as previously but to the three-body sheet $`\mathrm{\Pi }_3`$.
From ordinary potential scattering theory it is known that the poles of the scattering matrix corresponding to virtual or resonant states are located in the nonphysical energy domain. The continuation into this area may be performed by going around the branch point at threshold in order to get the virtual pole or directly from the upper half complex energy plane by crossing the real positive axis to reach the resonance pole. This continuation may clearly be done if the analytical expression for the scattering matrix is available. Otherwise, one has to use the dynamical equation written in terms of the real energies in order to continue it into the nonphysical area. In the former case, several recipes may be used (for a review of the methods see, e.g. ).
Turning to the three-body problem we would like to mention the method of analytical continuation based on a contour deformation technique. This method, developed in , was previously applied to the three-neutron problem and later to the study of the $`\mathrm{\Sigma }^{}nn`$\- and $`\mathrm{\Lambda }nn`$-interaction . In this paper we extend this technique to the $`\eta NN`$-interaction including the inelastic $`\pi NN`$-channel. Firstly, we would like to review briefly the basic details of the method. Let us consider the following Fredholm equation
$$F(p,W)=\underset{0}{\overset{\mathrm{}}{}}\frac{f(p^{},W)}{2mWp^2p^2(\stackrel{}{p}+\stackrel{}{p}^{})^2}F(p^{},W)๐p^{},$$
(37)
where the function $`f(p^{},W)`$ does not have any singularity in the relevant energy domain. For simplicity we consider three particles having equal masses $`m`$ and use nonrelativistic kinematics. The energy $`W`$ is the total kinetic energy. The physical region is determined in the usual fashion: $`\mathrm{}eW>0`$, $`\mathrm{}mW=\epsilon +0`$. Because of the singularities of the kernel for real $`W`$ one can not directly continue the equation down through the cut $`0W<\mathrm{}`$. In this case, shifting the integration path for $`p^{}`$ to the position $`C`$ in Fig. 4 (the angle $`\theta `$ is arbitrarily taken to be $`\pi /4`$) we can cross the real energy axis and enter the area on the nonphysical sheet. Here the variable $`p`$ is also taken on the contour $`C`$. The parabolic borderline separates the analytical domain from the forbidden area where the denominator in (37) may vanish for some values of $`p`$ and $`p^{}`$. If one uses relativistic kinematics the permissible area slightly narrows. The continuation procedure is of course justified when the contour being deformed does not hit any singularities of the integration kernel. It must be noted that the method allows one to uncover only the certain part of the nonphysical sheet. However, varying suitably the contour parameters $`(a,\theta )`$, one can get the major part of the nonphysical domain. We note also that in the present calculation the procedure described above was applied directly to Eq. (36), since all singularities of the kernel are contained in the Fredholm determinant $`D(W)`$.
### A Continuation into the lower half of the sheet $`\mathrm{\Pi }_1`$
From the diagrams shown in Fig. 3 it is clear that the lower half-plane of the sheet $`\mathrm{\Pi }_1`$ in Fig. 2 is at once the lower half of the three-body nonphysical sheet for the $`\pi NN`$ channel reached from above by crossing the cut between $`\pi NN`$ and $`\eta d`$ thresholds. This area is indicated as (II) in Fig. 3. Therefore, when continuing into this domain, the logarithmic singularities of the three-body propagator in $`V_\pi (p,p^{},W)`$ present an obstacle. A slight shift of the integration path (analogously to the pattern in Fig. 4) into the fourth quadrant allows one to reach the area shown in Fig. 5. Some calculational problems within this method may occur when the energy $`W`$ approaches the real axis above the $`\eta NN`$ three-particle threshold since the analyticity domain degenerates into part of a straight line. In this limit one has to handle the numerics with greater accuracy.
### B Continuation into the sheet $`\mathrm{\Pi }_2`$
In order to continue Eq. (36) into the sheet $`\mathrm{\Pi }_2`$, we adopt the following procedure. Entering into this sheet is accompanied by the movement of the pole $`p_0`$ of the propagator $`\tau _d`$ into the fourth quadrant of $`p^{}`$. This requires the contour deformation as is shown on Fig. 6. In restoring the integration path to its original position, we pick up the new term in the potential $`V_d(p,p^{},W)`$ corresponding to the residue of the integrand in (31) at $`q=p_0`$
$`V_d(p,p^{},W)`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^2dq}{2\omega _\eta (q)}}V_N(p,q,W)\tau _d\left(W\omega _\eta (q)q^2/4M_N\right)V_N(p^{},q,W)`$ (39)
$`+i{\displaystyle \frac{g_d\beta _d^{3/2}}{2\sqrt{\pi }}}{\displaystyle \frac{p_0\sqrt{\frac{|ฯต_d|}{M_N}}}{\omega _\eta (p_0)+2M_N}}V_N(p,p_0,W)V_N(p^{},p_0,W),`$
where
$$p_0=2\sqrt{M_N}\left\{W+\right|ฯต_d|\sqrt{m_\eta ^2+4M_N(W+|ฯต_d|M_N)}\}^{1/2},$$
(40)
(compare with (31)). In this manner we may cross the two-body cut in the interval $`W_{\eta d}<W<W_{\eta NN}`$ and enter into the lower half-plane of the sheet $`\mathrm{\Pi }_2`$ (transition $`22^{}`$ in the notation of Fig. 2). The new domain of analyticity is shown in Fig. 6.
Continuation into the upper half-plane of the sheet $`\mathrm{\Pi }_2`$ (transition $`11^{}`$) may be performed in a like manner. In this case, the pole $`p_0`$ crosses the integration contour when moving from below into the first quadrant of the $`p^{}`$-plane.
### C Continuation into the sheet $`\mathrm{\Pi }_3`$
For the continuation into the sheet $`\mathrm{\Pi }_3`$ from the region above the segment $`W_{\eta NN}W<\mathrm{}`$ of the real energy axis, we use the same technique as was outlined in the case A. In order to reach the domain below the $`\eta NN`$ threshold it is necessary to take $`\theta >\pi /4`$. In this case, however, one has to avoid the singularities on the imaginary $`p^{}`$-axis which come from the regularization form factors as well as from the factors of the type $`1/\sqrt{2\omega _j(p)}`$ in the integrands in (30). Therefore, the rotation angle $`\theta `$ must be restricted to values less than $`\pi /2`$. Thereby, we also do not encounter any problem with the poles of $`\tau _d`$ and $`\tau _N^{}`$. Choosing the mesh points from the deformed contour $`C`$, the three-body energy $`W`$ can be taken to pass above the three-body threshold $`W_{\eta NN}`$ and down through the cut into the lower half-plane of the sheet $`\mathrm{\Pi }_3`$ as is shown in Fig. 7 (transition $`44^{}`$ in Fig. 2).
The analytical continuation into the upper half-plane of the sheet $`\mathrm{\Pi }_3`$ is easily possible if the pion-exchange potential is ignored. Otherwise the logarithmic singularities from $`V_\pi `$ bar the continuation path. For this reason, when entering this area from the lower half-plane of $`\mathrm{\Pi }_1`$, we will switch off the pion-exchange forces. In this case, in order to make the transition $`33^{}`$ in the notation of Fig. 2, one just has to shift the contour into the first quadrant of the integration variable. Taking into account the smallness of the contribution from $`V_\pi `$ as denoted above, we believe that neglection of the pion-exchange potential does not spoil significantly the quality of our results.
## V Results and discussion
In order to obtain a better insight into the $`\eta NN`$ low-energy dynamics, we have investigated the trajectories drawn by the $`S`$-matrix poles on the Riemann surface when varying the interaction parameters. In doing this, we are guided by the following consideration. We presume that the pole trajectory, on which one expects to find a resonance or virtual three-body state closest to threshold, will also contain the deepest bound state eigenvalue. In this regard, we artificially enhanced the strength of the $`\eta NN`$-forces until the first bound state (i.e., the pole on the sheet $`\mathrm{\Pi }_1`$) appears. Then we have approached the actual physical situation by weakening the interaction strength to the value determined by the parameters given in Sect. III. Following this variation, the pole moves on the Riemann surface and finally arrives at several positions, developing in this way the bound (virtual) or resonant three-body state. Clearly, it may also appear on a Riemann sheet far removed from the physical region and thus will not influence the real scattering processes.
Turning back to the $`\eta NN`$-system, we take as a varying parameter the coupling constant $`g_N^{}^{(\eta )}`$. As its physical value we consider that presented in (22). We will start with the $`(1^{};0)`$ channel, the quasideuteron case. Before proceeding further, in order to see how the pole may behave when approaching the threshold region, we will ignore for the moment being the $`\pi NN`$-channel by setting $`g_N^{}^{(\pi )}`$=0. Then we took $`g_N^{}^{(\eta )}`$=4 and found the bound state pole at $`W_{pol}=W_{\eta NN}8.08`$ MeV. As expected, the weakening of $`g_N^{}^{(\eta )}`$ implies the motion of the pole towards the threshold region. At $`g_N^{}^{(\eta )}`$=2.5 it overtakes the $`\eta d`$ two-body threshold and passes into the two-body sheet $`\mathrm{\Pi }_2`$ producing a virtual $`\eta d`$ state. Further weakening of the $`\eta N`$-attraction pulls the pole back along the negative real axis away from the physical region. We would like to note that this shape of the trajectory is what we can expect naively from ordinary two-body potential scattering. Owing to the absence of the centrifugal barrier, the $`s`$-wave pole develops a virtual state and not a resonance when passing into the nonphysical sheet (here we do not touch upon exotic cases like the multitude of $`s`$-wave resonances in a deep square well ). We may expect to encounter the same pole behaviour in a three-body case since our physical situation is in effect the two-particle interaction with one being complex.
Now including again the $`\pi NN`$ channel, we see that its effect is to shift the starting position of the pole downwards from the real axis and thus to transform the real $`\eta NN`$ bound state into the $`\pi NN`$ resonance (or, what amounts to the same, the $`\eta NN`$ bound state with a finite lifetime). Then decreasing $`g_N^{}^{(\eta )}`$ and allowing the eigenvalue $`W_{pol}`$ to pass the $`\eta NN`$ threshold, we reach the situation illustrated in Fig. 8, where some pole trajectories corresponding to the different values of $`g_N^{}^{(\pi )}`$ are shown. One can see that the widths of the hypothetical $`\eta NN`$ bound states are sufficiently less than that of the $`S_{11}`$(1535) resonance in this energy region. This is due partially to the fact that $`\eta NN`$-system spends a large fraction of time in the $`\eta (NN)`$ state (not in $`N(N^{})`$), which in this region has no direct coupling with any configuration in the continuum. This effect explains also the anomalous decrease of the $`\pi NN`$ resonance width when approaching the $`\eta `$-production threshold. As one can see from Fig. 8, with increasing $`g_N^{}^{(\pi )}`$ the trajectories are shifted downwards and the point where the pole meets the real energy axis moves to the right. For the physical value $`g_N^{}^{(\pi )}`$=1.5 this point is located at $`WW_{\eta NN}`$+1.51 MeV. The trajectory crosses the real axis above the three-body threshold and passes to the three-body sheet $`\mathrm{\Pi }_3`$ (dashed curve with open triangles in Fig. 8). If one go around the three-body threshold $`W_{\eta NN}`$ and enters the upper half plane of the two-body sheet $`\mathrm{\Pi }_2`$ (dotted line in Fig. 8), one finds the extension of the trajectory also into this sheet (solid curve with filled circles). We recall that we neglected the pion-exchange potential when calculating the trajectory passing into the sheet $`\mathrm{\Pi }_3`$ (see Sect. IV C). For this reason the parts of the last two trajectories lying on the sheet $`\mathrm{\Pi }_1`$ are not identical. We see however that the difference is not significant. The pole positions corresponding to $`g_N^{}^{(\eta )}`$=2.0 are $`W_{pol}=W_{\eta NN}(3.00i13.67)`$ MeV on the sheet $`\mathrm{\Pi }_3`$ and $`W_{pol}=W_{\eta d}(4.39i7.22)`$ MeV on the sheet $`\mathrm{\Pi }_2`$.
After these findings we have carried out a careful search for poles on the lower half-plane of the sheets $`\mathrm{\Pi }_2`$ and $`\mathrm{\Pi }_3`$. Of course, we were interested primarily in the energy area just beyond the real axis where possible resonant states may occur. Our conclusion is that no poles appear in the physically interesting domain at least for reasonable values of the interaction parameters. As for the presence of a possible resonance pole on the three-body sheet $`\mathrm{\Pi }_3`$, one can assume that it may be found beyond the complex unitarity cut generated by the $`NN^{}`$ break-up singularity in $`\tau _N^{}`$. In any case, this pole, if it exists, is far removed from the relevant energy region and will hardly influence the near threshold $`\eta NN`$ interaction. Thus within our model, the $`S`$-matrix poles nearest to the physical region are situated on the upper half plane of the sheets $`\mathrm{\Pi }_2`$ and $`\mathrm{\Pi }_3`$ and produce virtual rather than resonant states in the $`\eta NN`$-system.
Now we will consider the $`(J^\pi ;T)=(0^{};1)`$ channel. We make the same variation of the $`N^{}`$ coupling constants and obtain the trajectories depicted in Fig. 9. Switching off the $`\pi NN`$ channel, we found the corresponding trajectory moving around the $`\eta NN`$ threshold and ending in the third quadrant of the variable $`WW_{\eta NN}`$ on the sheet $`\mathrm{\Pi }_3`$. The similar pole behaviour was noted also for the $`\mathrm{\Sigma }^{}nn`$ and $`\mathrm{\Lambda }nn`$ $`s`$-wave configurations in Ref. . Increasing $`g_N^{}^{(\pi )}`$ and varying $`g_N^{}^{(\eta )}`$ from 4 to 2, we find the general picture to be qualitatively the same as in the quasideuteron case. One notes a visible shift of the trajectories towards the higher energies, which is most likely explained by the weakening of the $`NN`$-interaction in the singlet $`{}_{}{}^{1}S_{0}^{}`$ state in relation to the more attractive $`{}_{}{}^{3}S_{1}^{}`$ state. The trajectory corresponding to $`g_N^{}^{(\pi )}`$=1.5 brings the final position of the pole to the point $`W_{pol}=W_{\eta NN}+(14.06+i17.19)`$ MeV.
What are the physical consequences that we can extract from the results above? First, as we just have shown, it is unlikely that a low-energy resonance will be found in the $`(0^{};1)`$ or $`(1^{};0)`$ channels. On the other hand, the virtual poles on the sheets $`\mathrm{\Pi }_2`$ and $`\mathrm{\Pi }_3`$ may influence the value of the scattering amplitudes through its proximity to the threshold region. Thus the next logical step in this direction would be to calculate, e.g., the $`\eta d`$-scattering near threshold. This problem, when treated within a three-body approach, is rather complex in itself and is beyond the scope of the present paper. The corresponding investigations for the elastic and inelastic $`\eta d`$-scattering will be presented elsewhere (see also ). Here we adopt the approximate formula for the $`s`$-wave elastic two-body scattering, when the amplitude has a virtual pole near zero energy (see e.g. )
$$\sigma _{\eta d}=\frac{4\pi }{|p_{pol}p|^2},$$
(41)
where $`p`$ denotes the c.m. $`\eta d`$ momentum and $`p_{pol}`$ its pole position, determined as
$$p_{pol}=\sqrt{\frac{2M_dm_\eta }{M_d+m_\eta }(W_{pol}M_dm_\eta )},\mathrm{}mp_{pol}<0,$$
(42)
with $`M_d`$ being the deuteron mass. In Fig. 10 we show the result obtained for $`W_{pol}=W_{\eta d}(4.39i7.22)`$ MeV. One sees a strong enhancement of the cross section as we approach the threshold energy, an effect which is typical when the system possesses a weak virtual state. Therefore we conclude that in the region of a few MeV above threshold an anomalous resonance-like behaviour can in general be expected in the cross section of reactions with $`\eta NN`$-channels in the final state.
Finally we must note that our results are in contrast to those of Shevchenko et al. who claimed the existence of $`s`$-wave $`\eta d`$-resonances. They used the on-shell solution of the nonrelativistic three-body equation and found the counterclockwise rotation of the $`\eta d`$ scattering amplitude in the Argand plot. The reason for this disagreement is not completely clear to us. On the other hand, we would like to mention the qualitative agreement of our results with those presented in where a similar behaviour of the $`\eta d`$-scattering cross section in the low-energy region was found.
## VI Summary and conclusion
In order to answer the question whether there might exist a low-energy resonance or virtual $`\eta NN`$-state, we have studied the typical pole trajectories $`W_{pol}(g_N^{}^{(\eta )})`$ for the $`\eta NN`$-system in $`L=0`$ states. Decreasing the $`\eta N`$-interaction parameter $`g_N^{}^{(\eta )}`$, the pole overtakes the three-body threshold and moves into the upper half plane of the sheets $`\mathrm{\Pi }_2`$ and $`\mathrm{\Pi }_3`$ adjacent to the lower rim of the two- and three-body unitarity cuts on the physical sheet $`\mathrm{\Pi }_1`$. Since these areas are not directly connected with the physical one, we have to conclude that no three-body resonances can occupy these trajectories. Moreover, we can assume that the very pattern of the pole trajectories forbids a low-energy $`\eta NN`$ resonance independent of the actual values of interaction parameters. We expect this feature to be valid in general for ($`L=0`$)-$`\eta NN`$ configurations, at least within the model of a type presented here. On the other hand, our results point to a possible explanation of the strong enhancement of the $`\eta `$-production cross section near threshold. This effect may be assigned not to a resonance but to virtual $`\eta NN`$-states, indeed generated by the poles on the sheets $`\mathrm{\Pi }_2`$ and $`\mathrm{\Pi }_3`$.
We would like to emphasize that our conclusion is based essentially on the model of a pure $`s`$-wave interaction. The typical feature of an $`s`$-wave interaction known from the familiar two-body problem is to develop a virtual rather than a resonance state. In this context, it will be of interest to enrich the present study by including $`p`$-waves in the partial wave decomposition (25). It must, however, be kept in mind that $`p`$-waves are energetically far less favourable in a low-energy regime. Therefore, we think that even though a pole may come sufficiently close to threshold, quite a strong enhancement factor from a $`p`$-wave contribution would be needed in order to make a possible $`p`$-wave resonance visible in the cross section.
Another aspect, left out in the present work, is the so-called direct $`NN^{}`$-interaction which is not automatically accounted for in the conventional three-body approach. Due to the short-range character of this interaction, associated with its static nature, the corresponding terms are expected to give only a small contribution when compared to the full scattering amplitude (with respect to the role of these terms in $`\mathrm{\Delta }N`$-interaction see e.g. ). However, it should be remembered, that three-body low-energy dynamics is known to be rather sensitive to the model variation. Thus even a small perturbation may change some of the present results.
## ACKNOWLEDGMENTS
One of the authors (A. F.) is grateful to the theory group of the Institut fรผr Kernphysik at the Johannes Gutenberg-Universitรคt Mainz for the very kind hospitality. Special thanks goes to Dr. Michael Schwamb for many fruitful discussions.
## Appendix
For the driving terms in (30) and (33) one finds the following analytic expressions:
(a) meson exchange potential $`(j=\pi ,\eta )`$:
$`V_j(p,p^{},W)=`$ $`{\displaystyle \frac{g_N^{}^{(j)2}\beta _N^{}^{(j)4}}{16\pi }}\left({\displaystyle \frac{M_N+m_j}{M_N}}\right)^2{\displaystyle \frac{1}{pp^{}(\alpha _1^2+c^2)(\alpha _2^2+c^2)}}`$ (A2)
$`\times \{\mathrm{ln}{\displaystyle \frac{ca_{}}{ca_+}}[{\displaystyle \frac{\alpha _2^2+c^2}{2(\alpha _1^2\alpha _2^2)}}(2\mathrm{ln}{\displaystyle \frac{|A_1^+|}{|A_1^{}|}}i{\displaystyle \frac{c}{\alpha _1}}\mathrm{ln}{\displaystyle \frac{A_1^+A_1^{}}{(A_1^+A_1^{})^{}}})+(12)]\},`$
where
$`c`$ $`=`$ $`WE_N(p)E_N(p^{}),`$ (A3)
$`a_\pm ^2`$ $`=`$ $`m_j^2+(p\pm p^{})^2,`$ (A4)
$`\alpha _1^2`$ $`=`$ $`{\displaystyle \frac{M_N+m_j}{M_N}}\beta _N^{}^{(j)2}m_j^2+{\displaystyle \frac{m_j}{M_N}}p^{\mathrm{\hspace{0.17em}2}}{\displaystyle \frac{m_j}{M_N+m_j}}p^2,`$ (A5)
$`\alpha _2^2`$ $`=`$ $`{\displaystyle \frac{M_N+m_j}{M_N}}\beta _N^{}^{(j)2}m_j^2+{\displaystyle \frac{m_j}{M_N}}p^2{\displaystyle \frac{m_j}{M_N+m_j}}p^{\mathrm{\hspace{0.17em}2}},`$ (A6)
and $`A_k^\pm =\alpha _k\pm ia_\pm `$ for $`k=1,2`$. If $`\alpha _1=\alpha _2=\alpha `$, then with $`A^\pm =\alpha \pm ia_\pm `$
$`V_j(p,p^{},W)`$ $`=`$ $`{\displaystyle \frac{g_N^{}^{(j)2}\beta _N^{}^{(j)4}}{16\pi }}\left({\displaystyle \frac{M_N+m_j}{M_N}}\right)^2{\displaystyle \frac{1}{pp^{}(\alpha _1^2+c^2)^2}}`$ (A8)
$`\times \left\{\mathrm{ln}\left({\displaystyle \frac{ca_{}}{ca_+}}{\displaystyle \frac{|A_1^+|}{|A_1^{}|}}\right)+ic{\displaystyle \frac{3\alpha ^2+c^2}{4\alpha ^3}}\mathrm{ln}{\displaystyle \frac{A^+A^{}}{(A^+A^{})^{}}}+2\mathrm{}e\left[\left(i+{\displaystyle \frac{c}{\alpha }}\right){\displaystyle \frac{a_{}a_+}{A^+A^{}}}\right]\right\}.`$
(b) nucleon exchange potential:
$`V_N(p,p^{},W)`$ $`=`$ $`{\displaystyle \frac{g_N^{}^{(\eta )}g_d\beta _N^{}^{(\eta )2}\beta _d^2}{16\pi pp^{}}}{\displaystyle \frac{M_N(M_N+m_\eta )}{m_\eta }}{\displaystyle \frac{1}{(a_2a_1)(a_2+c)(a_1+c)}}`$ (A10)
$`\times \left\{(a_2+c)\mathrm{ln}{\displaystyle \frac{a_1+pp^{}}{a_1pp^{}}}(a_1+c)\mathrm{ln}{\displaystyle \frac{a_2+pp^{}}{a_2pp^{}}}(a_1a_2)\mathrm{ln}{\displaystyle \frac{c+pp^{}}{cpp^{}}}\right\},`$
where
$`a_1`$ $`=`$ $`{\displaystyle \frac{M_N+m_\eta }{2m_\eta }}\left(\beta _N^{}^{(\eta )2}+p^{\mathrm{\hspace{0.17em}2}}\right)+{\displaystyle \frac{m_\eta }{2(M_N+m_\eta )}}p^2,`$ (A11)
$`a_2`$ $`=`$ $`\beta _d^2+p^2+p^{\mathrm{\hspace{0.17em}2}}/4,`$ (A12)
$`c`$ $`=`$ $`M_N(WE_N(p)\omega _\eta (p^{})M_N)(p^2+p^{\mathrm{\hspace{0.17em}2}})/2.`$ (A13)
|
warning/0006/math0006208.html
|
ar5iv
|
text
|
# Towards the ample cone of ๐ฬ_{๐,๐}
## ยง0 Introduction and Statement of Results
The moduli space of stable curves is among the most studied objects in algebraic geometry. Nonetheless, its birational geometry remains largely a mystery and most Mori theoretic problems in the area are entirely open. Here we consider one of the most basic:
###### 0.1 Question (Mumford)
What are the ample divisors on $`\overline{M}_{g,n}`$?
There is a stratification of $`\overline{M}_{g,n}`$ by topological type where the codimension $`k`$ strata are the irreducible components of the locus parameterizing pointed curves with at least $`k`$ singular points. An ample divisor must intersect any one dimensional stratum positively. It is thus natural to consider the following conjecture.
###### 0.2 Conjecture: $F\_1(\mgn)$
A divisor on $`\overline{M}_{g,n}`$ is ample iff it has positive intersection with all $`1`$-dimensional strata. In other words, any effective curve in $`\overline{M}_{g,n}`$ is numerically equivalent to an effective combination of $`1`$-strata. Said differently still, every extremal ray of the Mori cone of effective curves $`\overline{NE}_1(\overline{M}_{g,n})`$ is generated by a one dimensional stratum.
$`F_1(\overline{M}_{0,n})`$ was previously conjectured by Fulton, whence our choice of notation. Special cases of $`F_1(\overline{M}_{0,n})`$ are known (see \[KeelMcKernan96\], ยง7) but the general case remains open. The central observation of this work is that the case of higher genus is no harder. Indeed, $`\overline{NE}_1(\overline{M}_{g,n})`$ is naturally a quotient of $`\overline{NE}_1(\overline{M}_{0,2g+n})\times ^0`$ and Conjecture (0.2) is true for all $`g`$ iff it is true for $`g=0`$: see (0.3 -0.5). Further, the second formulation of conjecture (0.2) holds except possibly for very degenerate families in $`\overline{M}_{g,n}`$: see (0.6). Moreover, we are able to give strong results on the contractions of $`\overline{M}_{g,n}`$ (i.e. morphisms with geometrically connected fibers from it to other projective varieties). For example we show that for $`g2`$ the only fibrations of $`\overline{M}_{g,n}`$ are compositions of a tautological fibration, given by dropping points, with a birational morphism, see (0.9-0.11). Here are precise statements of our results:
We work over an algebraically closed field of any characteristic other than $`2`$ (this last assumption is used only at the end of the proof of (3.5)).
By the locus of flag curves, we shall mean the image $`\overline{F}_{g,n}`$ of the morphism $`f:\overline{M}_{0,g+n}/S_g\overline{M}_{g,n}`$ induced by gluing on $`g`$ copies of the pointed rational elliptic curve at $`g`$ points (which the symmetric group $`S_gS_{g+n}`$ permutes). The map $`f`$ is the normalization of $`\overline{F}_{g,n}`$.
###### 0.3 Theorem
$`g1`$. A divisor $`D\mathrm{Pic}(\overline{M}_{g,n})`$ is nef iff $`D`$ has non-negative intersection with all the one dimensional strata and $`D|_{\overline{F}_{g,n}}`$ is nef. In particular, $`F_1(\overline{M}_{0,g+n}/S_g)`$ implies $`F_1(\overline{M}_{g,n})`$.
Using the results of \[KeelMcKernan96\], Theorem (0.3) has the following consequence.
###### 0.4 Corollary
$`F_1(\overline{M}_{g,n})`$ holds in characteristic zero as long as $`g+n7`$ and when $`n=0`$ for $`g11`$.
For $`g4`$, $`F_1(\overline{M}_g)`$ was obtained previously by Faber.
We call $`E\overline{M}_{g,n}`$ ($`g1`$) a family of elliptic tails if it is the image of the map $`\overline{M}_{1,1}\overline{M}_{g,n}`$ obtained by attaching a fixed $`n+1`$-pointed curve of genus $`g1`$ to the moving pointed elliptic curve. Any two families of elliptic tails are numerically equivalent. In (2.2), we will see that all the $`1`$-strata except for $`E`$ are numerically equivalent to families of rational curves. Note that we abuse language and refer to a curve as rational if all of its irreducible components are rational. Thus it is immediate from (0.3) that the Mori cone is generated by $`E`$ together with curves in $`\overline{R}_{g,n}\overline{M}_{g,n}`$, the locus of rational curves. The locus $`\overline{R}_{g,n}`$ is itself (up to normalization) a quotient of $`\overline{M}_{0,2g+n}`$: By deformation theory $`\overline{R}_{g,n}`$ is the closure of the locus of irreducible $`g`$-nodal (necessarily rational) curves. The normalization of such a curve is a smooth rational curve with $`2g+n`$ marked points. Thus, if $`GS_{2g}`$ is the subgroup of permutations commuting with the product of $`g`$ transpositions $`(12)(34)\mathrm{}(2g\mathrm{1\hspace{0.17em}2}g)`$, the normalization of $`\overline{R}_{g,n}`$ is naturally identified with $`\overline{M}_{0,2g+n}/G`$. Thus (0.3) implies the following result:
###### 0.5 Corollary
$`D\mathrm{Pic}(\overline{M}_{g,n})`$ is nef iff $`D|_{\overline{R}_{g,n}E}`$ is nef. Equivalently, the natural map
$$\overline{NE}_1(\overline{M}_{0,2g+n}/GE)=\overline{NE}_1(\overline{M}_{0,2g+n}/G)\times ^0\overline{NE}_1(\overline{M}_{g,n})$$
is surjective.
In fact we prove the following strengthening of (0.3):
###### 0.6 Theorem
Let $`g1`$ and let $`N\overline{NE}_1(\overline{M}_{g,n})`$ be the subcone generated by the strata (1-5) of (2.2) for $`g3`$, by (1) and (3-5) for $`g=2`$, and by (1) and (5) for $`g=1`$. Then $`N`$ is the subcone generated by curves $`C\overline{M}_{g,n}`$ whose associated family of curves has no moving smooth rational components. Hence,
$$NE_1(\overline{M}_{g,n})=N+\overline{NE}_1(\overline{M}_{0,g+n}/S_g)$$
(See (1.2) for the meaning of moving component.) An analogous result for families whose general member has at most one singularity is given in \[Moriwaki00, A\].
Of course Theorem (0.6) implies (the second formulation) of Conjecture (0.2) for any family $`C\overline{M}_{g,n}`$ with no moving smooth rational component. In fact, $`F_1(\overline{M}_{0,7})`$ holds, and so Conjecture (0.2) holds as long as there is no moving smooth rational component containing at least $`8`$ distinguished (i.e. marked or singular) points.
Theorem (0.3) also has a converse which we state as follows.
###### 0.7 Theorem
Every nef line bundle on $`\overline{M}_{0,g+n}/S_g`$ is the pullback of a nef line bundle on $`\overline{M}_{g,n}`$ and $`F_1(\overline{M}_{0,g+n}/S_g)`$ is equivalent to $`F_1(\overline{M}_{g,n})`$. In particular, $`F_1(\overline{M}_g)`$ is equivalent to $`F_1(\overline{M}_{0,g}/S_g)`$ and $`F_1(\overline{M}_{1,n})`$ is equivalent to $`F_1(\overline{M}_{0,n+1})`$.
###### 0.8 Theorem
The Mori cone of $`\overline{F}_{g,n}`$ is a face of the Mori cone of $`\overline{M}_{g,n}`$: there is a nef divisor $`D`$ such that
$$D^{}\overline{NE}_1(\overline{M}_{g,n})=\overline{NE}_1(\overline{M}_{0,g+n}/S_g).$$
###### Remark Remark
An interesting problem is to determine whether or not $`F_1(\overline{M}_g)`$ for $`g1`$ implies $`F_1(\overline{M}_{0,n})`$ for all $`n3`$. For example, if one maps $`\overline{M}_{0,n}`$ into $`\overline{M}_g`$ by gluing one pointed curves of very different genera, then the pullback on the Picard groups will be surjective. However, it is not clear that this will be true for the nef cones.
We note that in a number of respects using $`\overline{F}_{g,n}`$ to reduce the general conjecture is more desirable than using $`\overline{R}_{g,n}`$. Analogues of Theorems (0.7) and (0.8) fail for $`\overline{R}_{g,n}`$, indicating that $`\overline{F}_{g,n}`$ is the more natural locus from the Mori theoretic point of view (e.g. one is led to consider $`\overline{F}_{g,n}`$ by purely combinatorial consideration of the Mori cone). Further the birational geometry of $`\overline{F}_{g,n}`$ is much simpler than that of $`\overline{R}_{g,n}`$. For example, when $`n=0`$, the Picard group of $`\overline{M}_{0,g}/S_g`$ has rank roughly $`g/2`$ and is freely generated by the boundary divisors, while $`\mathrm{Pic}(\overline{M}_{0,2g}/G)`$ has rank roughly $`g^2/2`$ and there is a relation among the boundary classes. Of greater geometric significance is the contrast between the cones of effective divisors: $`\overline{NE}^1(\overline{M}_{0,g}/S_g)`$ is simplicial and is generated by the boundary divisors (see \[KeelMcKernan96\]), while the structure of $`\overline{NE}^1(\overline{M}_{0,2g}/G)`$ is unclear and it is definitely not generated by boundary divisors (a counter-example is noted below).
It is straightforward to enumerate the possibilities for the one dimensional strata, and then express (0.2) as a conjectural description of the ample cone as an intersection of explicit half spaces. This description is given for $`n=0`$ in \[Faber96\] and for $`g=0`$ in \[KeelMcKernan96\]. We treat the general case in ยง2.
A fundamental problem in birational geometry is to study morphisms of a given variety to other varieties. In the projective category any such morphism is given by a semi-ample divisor โ i.e. a divisor such that the linear system of some positive multiple is basepoint free. A semi-ample divisor is necessarily nef. This implication is one of the main reasons for considering nef divisors. Correspondingly, one of the main reasons that (0.1) is interesting is its connection to:
###### Question
What are all the contractions (i.e. morphisms with geometrically connected fibers to projective varieties) of $`\overline{M}_{g,n}`$?
We have a number of results in this direction. Recall a definition from \[Keel99\]: For a nef divisor $`D`$ on a variety $`X`$ a subvariety $`ZX`$ is called $`D`$-exceptional if the $`D`$-degree of $`Z`$ is zero, or equivalently, $`D|_Z`$ is not big. The exceptional locus of $`D`$, denoted $`๐ผ(D)`$, is the union of the exceptional subvarieties. If $`D`$ is semi-ample and big, this is the exceptional locus for the associated birational map.
###### 0.9 Theorem
Let $`D\mathrm{Pic}(\overline{M}_{g,n})`$ be a nef divisor, $`g2`$ (resp. $`g=1`$). Either $`D`$ is the pull back of a nef divisor on $`\overline{M}_{g,n1}`$ via one of the tautological projections (resp. is the tensor product of pullbacks of nef divisors on $`\overline{M}_{1,S}`$ and $`\overline{M}_{1,S^c}`$ via the tautological projection for some subset $`SN`$) or $`D`$ is big and $`๐ผ(D)\overline{M}_{g,n}`$.
###### 0.10 Corollary
For $`g2`$ any fibration of $`\overline{M}_{g,n}`$ (to a projective variety) factors through a projection to some $`\overline{M}_{g,i}`$ ($`i<n`$), while $`\overline{M}_g`$ has no non-trivial fibrations.
The above corollary and its proof are part of the first authorโs Ph.D. thesis, \[Gibney00\], written under the direction of the second author.
###### 0.11 Corollary
$`g1`$. Let $`f:\overline{M}_{g,n}X`$ be a birational morphism to a projective variety. Then the exceptional locus of $`f`$ is contained in $`\overline{M}_{g,n}`$. In particular $`X`$ is again a compactification of $`M_{g,n}`$.
Note that our results give some support to the conjecture that the nef cone of $`\overline{M}_{g,n}`$ behaves as if it were described by Moriโs cone theorem (i.e. like the cone of a log Fano variety) with dual cone generated by finitely many rational curves and having every face contractible. This is surprising, since for example $`\overline{M}_{g,n}`$ is usually (e.g. for $`g24`$) of general type, and conjecturally (by Lang, see \[CHM97\]) no rational curve meets the interior of $`\overline{M}_{g,n}`$ for $`g2`$ and $`n`$ sufficiently big. Of course, our main results reduce Conjecture (0.2) to genus $`0`$, where from initial Mori theoretic considerations it is much more plausible.
We note that prior to this work very few nef (but not ample) line bundles on $`\overline{M}_{g,n}`$ were known. Myriad examples can be constructed using (0.3) as we indicate in ยง6.
The above results indicate that (regular) contractions give only a narrow view of the geometry of $`\overline{M}_{g,n}`$. For example, (0.10) gives that the only fibrations are (induced from) the tautological ones, and by (0.11), birational morphisms only affect the boundary. On the other hand the birational geometry should be very rich. For example, as is seen in (6.4), all but one of the elementary (i.e. relative Picard number one) extremal contractions of $`\overline{M}_g`$ are small (i.e. isomorphisms in codimension one), so there should be a wealth of flips. Also one expects in many cases to find interesting rational fibrations. For example, when $`g+1`$ has more than one factorization, the Brill-Noether divisor should give a positive dimensional linear system on $`\overline{M}_g`$ which is conjecturally (see \[HarrisMorrison98, 6.63\]) not big.
One consequence of (0.9) and the fact that $`๐ผ(\lambda )=\overline{M}_g`$ is that $`\overline{M}_g`$ is intrinsic.
###### 0.12 Corollary
Any automorphism of $`\overline{M}_g`$ must preserve the boundary.
In view of our results, $`F_1(\overline{M}_{0,n})`$ is obviously of compelling interest. It is natural to wonder:
###### 0.13 Question
If a divisor on $`\overline{M}_{0,n}`$ has non-negative intersection with all one dimensional strata, does it follow that the divisor is linearly equivalent to an effective combination of boundary divisors?
A positive answer to (0.13) would imply (0.2). The analogous property does hold on $`\overline{M}_{1,n}`$, (see (3.3)) but the question in that case is vastly simpler as the boundary divisors are linearly independent.
It is not true that every effective divisor on $`\overline{M}_{0,n}`$ is an effective sum of boundary divisors. In other words, $`F^1(\overline{M}_{0,n})`$ is false. In particular, the second author has shown (in support of the slope conjecture, \[HarrisMorrison98, 6.63\]) that the pullback of the Brill-Noether divisor from $`\overline{M}_g`$ to $`\overline{M}_{0,2g}`$ is a counter-example to $`F^1(\overline{M}_{0,2g+n}/G)`$ for any $`g3`$ with $`g+1`$ composite. In fact, he has shown that for $`g=3`$ the unique non-boundary component of this pullback gives a new vertex of $`\overline{NE}^1(\overline{M}_{0,6}/G)`$.
Question (0.13) can be formulated as an elementary combinatorial question of whether one explicit polyhedral cone is contained in another. As this restatement might be of interest to someone without knowledge of $`\overline{M}_g`$, we give this formulation:
###### 0.14 Combinatorial Formulation of (0.13)
Let $`๐`$ be the $``$-vector space that is spanned by symbols $`\delta _T`$ for each subset $`T\{1,2,\mathrm{},n\}`$ subject to the relations
Let $`N๐`$ be the set of elements $`b_T\delta _T`$ satisfying
$$b_{IJ}+b_{IK}+b_{IL}b_I+b_J+b_K+b_L,$$
for each partition of $`\{1,2,\mathrm{},n\}`$ into $`4`$ disjoint subsets $`I,J,K,L`$. Let $`E๐`$ be the affine hull of the $`\delta _T`$.
Question: Is $`NE`$?
In theory (0.14) can be checked, for a given $`n`$, by computer. For $`n6`$ we have done so, with considerable help from Maroung Zou, using the program Porta. Unfortunately the computational complexity is enormous, and beyond our machineโs capabilities already for $`n=7`$. These cases ($`n6`$) were proved previously (by hand) by Faber.
The remainder of the paper is organized as follows. In ยง1, we fix notation for various boundary divisors and gluing maps. In ยง2, we give a small set of generators for the cone spanned by one-dimensional strata (2.2) and inequalities which cut out the dual cone of divisors (2.1) โ we term each the Faber cone. The main result of ยง3 is (3.1) stating that a class nef on the boundary of $`\overline{M}_{g,n}`$ is nef. In ยง4 the results from the previous section are used to deduce (0.6) and hence (0.3). We also study the nefness and exceptional loci of certain divisors in order to deduce (0.7) and (0.8). The proof of (0.9) is contained in ยง5 following preparatory lemmas (5.1-5.3) dealing with divisors on $`C^n/G`$ where $`C`$ is a smooth curve with automorphism group $`G`$. In ยง6, we collect some more ad hoc results. In particular, we answer a question posed by Faber (6.2) and recover the classical ampleness result of Cornalba-Harris (6.3). Finally, ยง7 contains a geometric reformulation of $`F_1(\overline{M}_{0,n})`$ and a review of the evidence (to our minds considerable) for (0.2).
## Thanks
We have had interesting and fruitful discussions with various people on topics related to this paper, including X. Chen, F. Cukierman, J. Harris, B. Hassett, J. Kollรกr, A. Logan, J. McKernan, A. Moriwaki, R. Pandharipande, S. Popescu, W. Rulla, S. L. Tan, G. Ziegler, and M. Zou. J. Kollรกr gave us a counter-example to a theorem claimed in an earlier version of the paper.
Especially inspiring and enlightening were lengthy discussions with Carel Faber. In particular Faber showed us his proofs of $`F_1(\overline{M}_g)`$, $`g4`$, and helped us derive the results of ยง2.
## ยง1 Notation
For the most part we use standard notation for divisors, line bundles, and loci on $`\overline{M}_{g,n}`$. See e.g. \[Faber96\] and \[Faber97\]with the following possible exception. By $`\omega _i`$ on $`\overline{M}_{g,n}`$, $`g2`$, we mean the pullback of the relative dualizing sheaf of the universal curve over $`\overline{M}_g`$ by the projection given by dropping all but the $`i^{th}`$ point. Note this is not (for $`n3`$) the relative dualizing sheaf for the map to $`\overline{M}_{g,n1}`$ given by dropping the $`i^{th}`$ point (the symbol is used variously in the literature).
We note that the $``$-Picard group of $`\overline{M}_{g,n}`$ is the same in all characteristics, by \[Moriwaki01\].
We let $`N:=\{1,2,\mathrm{},n\}`$. The cardinality of a finite set $`S`$ is denoted $`|S|`$.
For a divisor $`D`$ on a variety $`X`$ and $`Y`$ a closed subset of $`X`$, we say that $`D`$ is nef outside $`Y`$, if $`DC0`$ for all irreducible curves $`CY`$.
To obtain a symmetric description of boundary divisors on $`\overline{M}_{0,n}`$, we use partitions $`๐ฏ=[T,T^c]`$ of $`N`$ into disjoint subsets, each with at least two elements, rather than subsets $`T`$. We write the corresponding boundary divisor as $`\delta _{0,๐ฏ}`$, or $`\delta _๐ฏ`$. For a partition $`Q`$ of a subset $`SN`$ we write $`๐ฏ>Q`$ provided the equivalence relation induced by $`Q`$ on $`S`$ refines that obtained by restricting $`๐ฏ`$ to $`S`$. We write $`Q=๐ฏ|_S`$ provided these two equivalence relations are the same.
We write $`B_k`$ for the sum of boundary divisors on $`\overline{M}_{0,n}/S_n`$ that are the image of any $`\delta _๐ฏ`$ with $`|T|=k`$ (thus $`B_k=B_{nk}`$). We will abuse notation by using the same expression for its inverse image (with reduced structure) on $`\overline{M}_{0,n}`$, $`_{๐ฏ,|T|=k}\delta _๐ฏ`$.
We make repeated use of the standard product decomposition for strata as a finite image of products of various $`\overline{M}_{g,n}`$. For precise details see \[Keel99, pg. 274\] and \[Faber97\]. To describe this decomposition for boundary divisors we use the notations
$$\stackrel{~}{\mathrm{\Delta }}_{i,S}:=\overline{M}_{i,S}\times \overline{M}_{gi,S^c}\mathrm{\Delta }_{i,S}$$
and
$$\stackrel{~}{\mathrm{\Delta }}_{\mathrm{irr}}:=\overline{M}_{g1,Npq}\mathrm{\Delta }_{\mathrm{irr}}.$$
We refer to $`\overline{M}_{i,S}`$ (and any analogous term for a higher codimension stratum) as a factor of the stratum.
The key fact about these maps used here is the
###### 1.1 Lemma
The pullback to $`\stackrel{~}{\mathrm{\Delta }}_{i,S}`$ of any line bundle is numerically equivalent to a tensor product of unique line bundles from the two factors. The given line bundle is nef on $`\mathrm{\Delta }_{i,S}`$ iff each of the line bundles on the factors is nef. Dually, let $`C`$ be any curve on the product, and $`C_i,C_{gi}`$ be its images on the two factors (with multiplicity for the pushforward of cycles) which we also view as curves in $`\overline{M}_{g,n}`$ by the usual device of gluing on a fixed curve. Then, $`C`$ and $`C_i+C_{gi}`$ are numerically equivalent.
###### Demonstration Proof
The initial statement implies the other statements, and follows from the explicit formulae of \[Faber97\].โ
Any curve $`E`$ in $`\overline{M}_{g,n}`$ induces a decomposition of the curves it parameterizes into a subcurve fixed in the family $`E`$ and a moving subcurve. Arguing inductively, (1.1) yields,
###### 1.2 Corollary
Every curve in $`\overline{M}_{g,n}`$ is numerically equivalent to an effective combination of curves whose moving subcurves are all generically irreducible.
## ยง2. The Faber cone
In this section we consider the subcone of $`\overline{NE}_1(\overline{M}_{g,n})`$ generated by one dimensional strata, and its dual. We refer to either as the Faber cone, appending of curves, or of divisors if confusion is possible. We will call a divisor $`F`$-nef if it lies in the Faber cone. Of course (0.2) is then the statement that $`F`$-nef implies nef, or equivalently, the Mori and Faber cones of $`\overline{M}_{g,n}`$ are the same.
The next result describes the Faber cone of divisors as an intersection of half spaces. In order to give a symmetric description of the cone we write $`\delta _{0,\{i\}}`$ for $`\psi _i`$ in $`\mathrm{Pic}(\overline{M}_{g,n})`$ (so $`\delta _{0.I}`$ is defined whenever $`|I|1`$).
###### 2.1 Theorem
Consider the divisor
$$D=a\lambda b_{\mathrm{irr}}\delta _{\mathrm{irr}}\underset{\frac{}{[g/2]i0IN|I|1\text{ for }i=0}}{}b_{i,I}\delta _{i,I}$$
on $`\overline{M}_{g,n}`$ (with the convention that we omit for a given $`g,n`$ any terms for which the corresponding boundary divisor does not exist). Consider the inequalities
where $`b_{i,I}`$ is defined to be $`b_{gi,I^c}`$ for $`i>[g/2]`$.
For $`g3`$, $`D`$ has non-negative intersection with all $`1`$-dimensional strata iff each of the above inequalities holds.
For $`g=2`$, $`D`$ has non-negative intersection will all $`1`$-dimensional strata iff (1) and (3-6) hold.
For $`g=1`$, $`D`$ has non-negative intersection with all $`1`$-dimensional strata iff (1) and (5-6) hold.
For $`g=0`$, $`D`$ has non-negative intersection with all $`1`$-dimensional strata iff (6) holds.
###### Demonstration Proof
(2.2) below lists the numerical possibilities for a stratum. Each inequality above comes from standard intersection formulae, see e.g. \[Faber96\], \[Faber97\], by intersecting with the corresponding curve of (2.2). Intersecting with (2.2.4) gives the inequality $`2b_{\mathrm{irr}}b_{i+1,I}`$, which shifting indices and notation gives (2.1.4). โ
Theorem (2.2) gives a listing of numerical possibilities for one dimensional strata, giving explicit representatives for each numerical equivalence class. The parts in Theorems (2.1) and (2.2) correspond: that is, for each family $`X`$ listed in parts 2 to 6 of (2.2), the inequality which expresses the condition that a divisor $`D`$ given as in (2.1) meet $`X`$ non-negatively is given in the corresponding part of (2.1).
We obtain these just as in \[Faber96\], by defining a map $`\overline{M}_{0,4}\overline{M}_{g,n}`$ by attaching a fixed pointed curve in some prescribed way. For a subset $`IN`$, by a $`k+I`$ pointed curve we mean a $`k+|I|`$ pointed curve, where $`|I|`$ of the points are labeled by the elements of $`I`$.
###### 2.2 Theorem
Let $`X\overline{M}_{g,n}`$ be a one dimensional stratum. Then $`X`$ is either
or, numerically equivalent to the image of $`\overline{M}_{0,4}\overline{M}_{g,n}`$ defined by one of the attaching procedures 2-6 below.
Figure (2.3) shows schematic sketches of each of these 5 families numbered as in (2.2). The generic fiber is shown on the left of each sketch and the 3 special fibers (with, up to dual graph isomorphism, any multiplicities) are shown on the right. The bolder curves are the component(s) of the $`\overline{M}_{0,4}`$ piece. Boxes give the type (i.e. genus and marked point set) of each fixed component and of each node not of irreducible type.
In any of the families, if the curve sketched is not stable, we take the stabilization. Thus e.g. in (6) for $`n=4`$, $`g=0`$, the map $`\overline{M}_{0,4}\overline{M}_{0,4}`$ is the identity, while in (3) for $`g=2,n=0`$, $`i=j=0`$, the image of a generic point of $`\overline{M}_{0,4}`$ is the moduli point of an irreducible rational curve with two ordinary nodes.
Strata of type (6) play distinctly different roles, both geometrically and combinatorially, from those of type (1-5). Those of type (6) come from the flag locus, they are the only strata in genus $`0`$, and there is a face of $`\overline{NE}_1(\overline{M}_{0,n})`$ that contains exactly these strata, see (4.9). The cone generated by the strata (1-5) has a nice geometric meaning given in (0.6). Our proofs make no direct use of strata of type (6); in particular we will only directly use the inequalities (2.1.1 -2.1.5).
do{
Strata of type (2-6) are all (numerically equivalent to) curves lying in $`\overline{R}_{g,n}`$. Those of type (6) are distinguished geometrically by the fact that the curve corresponding to a general point has only disconnecting nodes, and algebraically by the fact that the corresponding inequality in (2.1) has more than $`3`$ (in fact $`7`$) terms.
###### Demonstration Proof of 2.2
The proof is analogous to that of the case of $`n=0`$ treated in \[Faber96\]. Here are details.
Let $`X\overline{M}_{g,n}`$ be a one dimensional stratum: $`X`$ is a finite image of a product of moduli spaces (one for each irreducible component of the curve $`C`$ corresponding to its general point). All but one of these spaces are zero dimensional, thus $`\overline{M}_{0,3}`$, and there is a distinguished one dimensional factor which is either $`\overline{M}_{1,1}`$ or $`\overline{M}_{0,4}`$. In the $`\overline{M}_{1,1}`$ case, $`X`$ is a family of elliptic tails and gives (2.1.1) exactly as in \[Faber96\].
Consider the $`\overline{M}_{0,4}`$ case. Let $`C`$ be the stable pointed curve corresponding to a general point of $`X`$ and $`EC`$ be the moving irreducible component (which is rational). Let $`h:\stackrel{~}{E}E`$ be the normalization and let $`M\stackrel{~}{E}`$ be the union of the following disjoint subsets: $`h^1(ZE)`$ for each connected component $`Z`$ of the closure of $`CE`$, $`h^1(p)`$ for each singular point $`p`$ of $`E`$, and $`h^1(p_i)`$ for each labeled point $`p_i`$ of $`C`$ on $`E`$. These subsets partition $`M`$ into equivalence classes.
We assign to each equivalence class $`๐ต`$ a triple $`(a(๐ต),S(๐ต),h(๐ต))`$: if $`๐ต=h^1(ZE)`$, then $`h`$ is the arithmetic genus of $`Z`$, $`S\{1,2,\mathrm{},n\}`$ is the collection of labeled points of $`C`$ lying on $`Z`$, and $`a`$ is the cardinality of $`ZE`$. $`t(h^1(p))=(2,\mathrm{},0)`$ for a singular point $`p`$ of $`E`$, and $`t(h^1(p_i))=(1,\{i\},0)`$ for a labeled point $`p_i`$ of $`C`$ on $`E`$. The numerical class of $`X`$ is determined by the collection of triples $`๐ต`$ by the standard intersection product formulae in, for example, \[Faber97\]. Note that $`_๐ต(h(๐ต)+a(๐ต)1)=g`$, that $`_๐ตS(๐ต)`$ is a partition of $`\{1,2,\mathrm{},n\}`$, and that $`_๐ตa(Z)=4`$.
Now (2-6) are obtained by enumerating the possibilities for the collection of $`a(Z)`$ which are $`\{4\}`$, $`\{3,1\}`$, $`\{2,2\}`$, $`\{2,1,1\}`$ and $`\{1,1,1,1\}`$. These correspond to (the obvious generalizations to pointed curves of) the families (B), (C), (D), (E), and (F) of \[Faber96\] and yield (2-6) by considering the possibilities for $`h(๐ต)`$ and $`S(๐ต)`$ subject to the above constraints. โ
Throughout the paper we will separate the $`\psi `$ and $`\delta `$ classes (because although they have similar combinatorial properties, they are very different geometrically, e.g. $`\psi _i`$ is nef, while the boundary divisors have very negative normal bundles). Thus, we will write a divisor as
$$\underset{i=1}{\overset{n}{}}c_i\psi _i+a\lambda b_{\mathrm{irr}}\delta _{\mathrm{irr}}\underset{\frac{}{SNn2|S|2}}{}b_{0,S}\delta _{0,S}\underset{[g/2]j1,SN}{}b_{i,S}\delta _{i,S}$$
$`2.4`$
where $`b_{0,\{i\}}`$ has become $`c_i`$. As examples, in this notation (2.1.5) with $`i=j=0,I=\{i\}`$ becomes
$$c_i+b_{0,J}b_{0,J\{i\}}$$
which in turn specializes to
$$c_i+c_jb_{0,\{i,j\}}$$
in case $`J=\{j\}`$ while (2.1.3) for $`(i,I)=(0,\{i\})`$ becomes $`c_i0`$.
For $`g3`$, the expression above is unique (i.e. the various tautological classes are linearly independent), but for smaller genera there are relations which are listed in \[ArbarelloCornalba98\]. For $`g=1`$, the boundary divisors are linearly independent and span the Picard group so we can assume above that $`a`$ and all $`c_i`$ are zero and, if we do so, the resulting expression is unique.
## ยง3. Nef on the boundary implies nef
In this section, we prove:
###### 3.1 Proposition
If $`g2`$, or $`g=1,n2`$, a divisor $`D\mathrm{Pic}(\overline{M}_{g,n})`$ is nef iff its restriction to $`\overline{M}_{g,n}`$ is nef.
###### 3.2 Lemma
$`g2`$. Let $`D\mathrm{Pic}(\overline{M}_{g,n})`$ be a divisor expressed as in (2.1). For $`[g/2]i1`$ let
$$b_i=\underset{SN}{\text{max }}b_{i,S}$$
and define $`A\mathrm{Pic}(\overline{M}_g)`$ by
$$A=a\lambda b_{\mathrm{irr}}\delta _{\mathrm{irr}}\underset{[g/2]i1}{}b_i\delta _i.$$
If the coefficients of $`D`$ satisfy one of the inequalities (2.1.1-2.1.5) for fixed indices $`i,j`$ and all subsets of $`N`$, then the coefficients of $`A`$ satisfy the analogous inequality. In the example of (2.1.5), $`b_i+b_jb_{i+j}`$ if $`b_{i,I}+b_{j,J}b_{i+j,IJ}`$ for all $`i`$ and $`j`$ between $`0`$ and $`g1`$ and for all subsets $`I,JN`$ with $`IJ=\mathrm{}`$.
###### Demonstration Proof
This is clear since the $`b_i`$ are defined by taking a maximum. For example, suppose $`b_{i+j}=b_{i+j,T}`$. Then, $`b_i+b_jb_{i,\mathrm{}}+b_{j,T}b_{i+j,T}=b_{i+j}`$.โ
Since all the strata are contained in $`\overline{M}_{g,n}`$, (3.1) will follow immediately from:
###### 3.3 Lemma
Let $`D\mathrm{Pic}(\overline{M}_{g,n})`$. If $`g3`$ \[resp. g=2; g=1\] and $`D`$ has non-negative intersection with all strata in (2.2) of types (1-5) \[resp. (1-2) and (4-5); (1) and (5)\] then $`D`$ is nef outside of $`\overline{M}_{g,n}`$, and furthermore if $`g=1`$ then $`D`$ is numerically equivalent to an effective sum of boundary divisors.
###### Demonstration Proof
We express $`D`$ as in (2.4). We consider first the case of $`g=1`$. We can as remarked above assume $`a=c_i=0`$ for all $`i`$. Then it is immediate from (2.1.1) and (2.1.5) (using the translations after (2.4)) that all the coefficients are non-positive, so $`D`$ is linearly equivalent to an effective sum of boundary divisors.
Now assume $`g2`$. Define $`b_i`$ and $`A`$ as in (3.2).
We use the relation
$$\psi _i=\omega _i+\underset{iS}{}\delta _{0,S}$$
(which, as in \[ArbarelloCornalba96\], is obtained easily from the definition by induction on $`n`$) to write
$$D=\underset{iN}{}c_i\omega _i+\left(\underset{\frac{}{SNn2|S|2}}{}\left(\left(\underset{iS}{}c_i\right)b_{0,S}\right)\delta _{0,S}\right)+\pi ^{}(A)+E$$
$`3.4`$
where $`\pi `$ is the forgetful map to $`\overline{M}_g`$ and $`E`$ is a combination of boundary divisors $`b_{i,S},i>0`$ โ an effective one since we take the maximum in forming $`b_i`$.
By repeated application of (2.1.5) as translated below (2.4), the coefficients of the $`\delta _{0,S}`$ in (3.4) are non-negative. By (2.1.3), $`c_i0`$. Since each $`\omega _i`$ is nef โ by, e.g. \[Keel99\] โ itโs enough to show $`A`$ is nef outside $`\overline{M}_g`$.
Suppose first $`g3`$. Then $`A`$ satisfies each of (2.1.1-2.1.5) by (3.2). Thus $`a10b_{\mathrm{irr}}0`$ and $`a5b_i`$, so the claim follows from (3.5) below. Now suppose $`g=2`$. By (3.2) and (2.2) $`a10b_{\mathrm{irr}}2b_{\mathrm{irr}}b_10`$ so the claim follows from \[Faber96, ยง2\]. โ
###### 3.5 Lemma
Let $`D=a\lambda b_i\delta _i`$ be a divisor on $`\overline{M}_g`$, and let $`g:Z\overline{M}_g`$ a morphism from a projective variety such that $`g(Z)\overline{M}_g`$. Assume $`g2`$. Consider the inequalities:
If (1-3) hold then $`g^{}(Z)`$ is a limit of effective Weil divisors and if strict inequality holds in each case and $`g`$ is generically finite then $`g^{}(Z)`$ is big.
###### Demonstration Proof
The final statement follows from the previous statements using $`12\lambda =\kappa +\delta `$ and the ampleness of $`\kappa `$.
Let $`\mathrm{\Gamma }`$ be the divisor of \[CornalbaHarris88, 4.4\]
$$(8g+4)\lambda g\delta _{\mathrm{irr}}2g\underset{i>0}{}\delta _i.$$
The inequalities imply $`D=s\lambda +c\mathrm{\Gamma }+E`$ where $`c,s0`$, and $`E`$ is an effective sum of boundary divisors. The pullback $`g^{}(\mathrm{\Gamma })`$ is a limit of effective Weil divisors by a slight generalization of \[CornalbaHarris88, 4.4\] (in characteristic zero we could apply \[Moriwaki98\] directly). When $`g(Z)`$ is contained in the hyperelliptic locus, their analysis applies without change, as long as the characteristic is not two. When $`g(Z)`$ is not contained in the hyperelliptic locus, one needs the generalization of their (2.9) to smooth $`T`$ of arbitrary dimension (with their $`e_L()`$ replaced by the analogous expression supported at codimension one points of $`T`$ and the conclusion being that the expression is a limit of effective Weil divisors, as in their (1.1)). Their proof extends after obvious modifications. โ
###### 3.6 Theorem
Let $`D\mathrm{Pic}(\overline{M}_{g,n})`$ be a divisor with non-negative intersection with all $`1`$-strata of type (2.2.1-2.2.5). Let $`C\overline{M}_{g,n}`$ be a complete one dimensional family of curves, whose generic member is a curve without moving rational components. Then $`DC0`$. If furthermore $`D|_{\overline{R}_{g,n}}`$ is nef, then $`D`$ is nef.
###### Demonstration Proof
We induce simultaneously on $`g,n`$. For $`g=0`$ or $`g=1,n=1`$ there is nothing to prove. So suppose $`g2`$, or $`g=1,n2`$.
Suppose first that $`D`$ is nef on all $`1`$-strata of type (1-5). Let $`C`$ be as in the statement. We will show that $`DC0`$. By (3.3) we may assume $`C\overline{M}_{g,n}`$. Suppose $`C\mathrm{\Delta }_{i,S}`$ for $`i>0`$. We apply induction using (1.1), following that notation. Itโs enough to show $`D_iC_i0`$. A stratum of type (1-5) on one of the factors yields (by the usual inclusion given by gluing on a fixed curve) a stratum of type (1-5) on $`\overline{M}_{g,n}`$, so $`D_i,C_i`$ satisfy the induction hypothesis. If $`C\mathrm{\Delta }_{\mathrm{irr}}`$ we can pass to $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{irr}}`$ and apply induction directly.
Now suppose that in addition that $`D|_{\overline{R}_{g,n}}`$ is nef. We use induction just as above (with the same notation) Note that we may assume that $`D_i`$ is nef on $`\overline{R}_{i,S}`$, since we may choose the fixed curve we glue on to be rational. โ
## ยง4 Reduction to the Flag locus
###### 4.1 Prop
Let $`D`$ be a divisor on $`\overline{M}_{g,n}`$, $`g1`$, satisfying the hypothesis of (3.3). Let $`X\overline{R}_{g,n}`$ be a stratum whose general member corresponds to a stable curve with no disconnecting nodes. Then $`D|_X`$ is linearly equivalent to an effective sum of boundary divisors and nef divisors.
###### 4.2 Corollary
Let $`C\overline{M}_{g,n}`$ be a complete one dimensional family of pointed curves whose general member has no moving smooth rational components. Let $`D`$ be a divisor on $`\overline{M}_{g,n}`$. If $`D`$ has non-negative intersection with all $`1`$-strata of type (2.2.1-2.2.5), then $`DC0`$.
###### Demonstration Proof
We use induction as in the proof of (3.6). By (1.2) we can assume that the general member of $`C`$ is irreducible, and by (3.6), rational. Now (4.1) applies. โ
###### 4.3 Corollary
$$NE_1(\overline{M}_{g,n})=N+\overline{NE}_1(\overline{M}_{0,g+n}/S_g)$$
where $`N`$ is the subcone generated by the strata (2.2.1-2.2.5).
###### Demonstration Proof
Immediate from (4.2), (1.1) and (1.2). โ
Of course (4.2) and (4.3) together contain (0.6), which in turn implies (0.3).
###### Demonstration Proof of (4.1)
First we reduce to the case of $`n=0`$. As in (3.4) we have
$$D=\underset{iN}{}c_i\omega _i+\pi ^{}(A)+G$$
with $`G`$ an effective combination of boundary divisors parameterizing degenerations with a disconnecting node. (In the genus $`1`$ case the sum is empty and $`A`$ is a multiple of $`\delta _{\mathrm{irr}}`$). So we can replace $`X`$ by its image under $`\pi `$. When $`g=1`$ this is a point and there is nothing to prove. When $`g=2`$ this image is either a point, or is one dimensional and defines inequality (2.1.4) (so the pullback of $`D`$ is nef). So we can assume $`n=0,g3`$.
Let $`C`$ be the stable curve corresponding to a general point of $`X`$, and $`E`$ a component of $`C`$. As in the proof of (2.2) we get from $`E`$ a map of some $`\overline{M}_{0,n}`$ into $`\overline{M}_g`$ and we want to see the pullback of $`D`$ is an effective sum of boundary components and nef divisors. As in the proof of (2.2), the $`n`$ points are divided into equivalence classes: the singular points of $`E`$ are divided into disjoint $`2`$-cycles, while the points where $`E`$ meets other irreducible components of $`C`$ are divided by the connected components of (the closure of) $`CE`$, each class with at least $`2`$ elements by our assumption that $`C`$ has no disconnecting nodes. If $`P`$ is the corresponding partition, then a class $`\delta _i`$, $`i>0`$ pulls back to a sum of irreducible divisors $`\delta _๐ฏ`$ with $`๐ฏ>P`$ (i.e. $`P`$ refining the partition $`[T,T^c]`$). We use the identity on $`\overline{M}_g`$
$$\delta _{\mathrm{irr}}+12\lambda =\kappa +\underset{i>0}{}\delta _i$$
and the fact that $`\lambda `$ is trivial on $`\overline{M}_{0,n}`$. If $`b_{\mathrm{irr}}0`$ then $`b_{\mathrm{irr}}=0`$ by (2.1.2) and so the result is clear from (2.1.4). So we may assume (by scaling) that $`b_{\mathrm{irr}}=1`$. Then $`D`$ pulls back to
$$\kappa +\underset{๐ฏ>P}{}e_T\delta _๐ฏ$$
with $`e_T1`$ by (2.1.4). Now apply (4.4) below. โ
###### 4.4 Lemma
Let $`P`$ be a partition of $`\{1,2,\mathrm{},n\}`$, such that each equivalence class contains at least two elements. The divisor
$$\kappa +\underset{๐ฏ>P}{}e_๐ฏ\delta _๐ฏ$$
$`4.5`$
on $`\overline{M}_{0,n}`$ is linearly equivalent to an effective sum of boundary divisors, if each $`e_๐ฏ1`$.
###### Remark Remark
Note that at least some condition on $`P`$ is required. E.g. $`\kappa \delta `$ is not an effective sum of boundary divisors, since as a symmetric expression the coefficient of $`B_2`$ is negative, and there are no symmetric relations.
###### Demonstration Proof of (4.4)
Expanding $`\kappa `$ in terms of boundary divisors (see e.g. \[KeelMcKernan96\]) we have
$$\underset{๐ฏ>P}{}(\frac{|T|(n|T|)}{n1}+e_๐ฏ1)\delta _๐ฏ+\underset{๐ฏP}{}(\frac{|T|(n|T|)}{n1}1)\delta _๐ฏ.$$
Note $`\frac{|T|(n|T|)}{n1}2`$ unless $`|T|=2`$, in which case it is $`2\frac{2}{n1}`$, or $`n=6`$ and $`|T|=3`$ in which case it is $`2\frac{1}{n1}`$. So the sum is nearly effective as it stands. We use relations in the Picard group to deal with the negative coefficients.
We assume first $`n7`$. If there are no $`2`$-cycles in $`P`$ our sum is already effective, so (after perhaps reordering) we can write $`P`$ as
$$(12)(34)\mathrm{}(2k\mathrm{1\; 2}k)$$
for some $`k1`$. Further we can assume either that $`2k=n`$, or we have a further $`t`$-cycle, $`t3`$, $`(2k+\mathrm{1\; 2}k+2\mathrm{}2k+t)`$. We use the $`k`$ relations
$$\underset{[\{i,i+1\},\{i+2,i+3\}]=๐ฏ|_{\{i,i+1,i+2,i+3\}}}{}\delta _๐ฏ=\underset{[\{i,i+2\},\{i+1,i+3\}]=๐ฏ|_{\{i,i+1,i+2,i+3\}}}{}\delta _๐ฏ$$
for $`i=1,3,5,\mathrm{},2k1`$ with the adjustment that if $`2k=n`$, then we replace the last by the analogous relation for $`[\{2k1,2k\},\{1,2\}]`$. By $`[\{i,i+1\},\{i+2,i+3\}]=๐ฏ|_{\{i,i+1,i+2,i+3\}}`$ we mean that the restriction of $`๐ฏ`$ to the $`4`$-element set is the given partition into disjoint $`2`$-cycles, i.e. either $`i,i+1T`$ and $`i+2,i+3T^c`$ or this holds with $`T`$ and $`T^c`$ reversed. Thus the above relations come directly from those in \[Keel92\] by adjusting notation.
Let $`L=R`$ be the relation among effective boundary divisors that one obtains by adding together the $`k`$ relations above (without canceling terms common to both sides). Let $`\delta _{\mathrm{irr}}`$ be the sum of the $`\delta _๐ฏ`$ with $`๐ฏP`$, and $`\delta _1`$ the sum of those with $`๐ฏ`$ one of the $`2`$-cycles of $`P`$ together with its complement. Note $`R`$ is supported on $`\delta _{\mathrm{irr}}`$ while $`L\delta _1`$ (here and throughout the paper inequalities between sums of Weil divisors stand for a system of inequalities, one for each component) and $`L2\delta _1`$ when $`n=2k`$.
Choose $`s>0`$ maximal so that
$$sR\underset{๐ฏP}{}(\frac{|T|(n|T|)}{n1}1)\delta _๐ฏ.$$
Thus (4.5) is linearly equivalent to
$$sL\frac{2}{n1}\delta _1+E$$
for $`E`$ an effective sum of boundary divisors.
Define coefficients $`\alpha _๐ฏ`$ so $`R=_๐ฏ\alpha _๐ฏ\delta _๐ฏ`$.
Since $`L\delta _1`$, it is enough to show that $`s\frac{2}{n1}`$, or equivalently that
$$\frac{2}{n1}\alpha _๐ฏ\frac{|T|(n|T|)}{n1}1$$
$`4.6`$
and for $`n=2k`$, $`\frac{\alpha _๐ฏ}{n1}\frac{|T|(n|T|)}{n1}1`$ is sufficient. Suppose first $`|T|=2`$. Note that $`\delta _{0,T}`$ can occur (with non-zero coefficient) on the right hand side of at most one of the $`k`$ relations. Thus $`\alpha _๐ฏ1`$ and (4.6) follows.
Suppose now $`|T|3`$. Then (4.6) holds as long as $`2\alpha _๐ฏn1`$. Since $`\alpha _๐ฏk`$, we can assume $`2k=n`$. Then we only need $`\alpha _๐ฏ2k1`$, which holds as $`\alpha _๐ฏk`$.
Now we consider the cases of $`n6`$. If $`P`$ has an $`n`$-cycle, the result is obvious ($`๐ฏ>P`$ is then impossible). Up to reordering the possibilities are $`(12)(34)`$, $`(12)(345)`$, $`(12)(34)(56)`$ and $`(123)(456)`$. In the first case (4.5) is nef (and we are working on $`\overline{M}_{0,4}=^1`$). In the second case the above argument applies without change. In the third we can combine the above argument for the cases $`n=2k`$ and $`|T|=2`$. Finally consider $`P=(123)(456)`$. Thus $`๐ฏ>P`$ implies $`\delta _๐ฏ=\delta _{0,\{1,2,3,\}}`$ We use the relation
$$\underset{[\{1,2\},\{5,6\}]=๐ฏ|_{\{1,2,5,6\}}}{}\delta _๐ฏ=\underset{[\{1,5\},\{2,6\}]=๐ฏ|_{\{1,2,5,6\}}}{}\delta _๐ฏ$$
to show (4.5) is effective as above. โ
###### 4.7 Theorem
Consider the line bundle
$$D=a\lambda +\underset{i=1}{\overset{i=n}{}}(g+n1)\psi _ib_{\mathrm{irr}}\delta _{\mathrm{irr}}\left(g+n(i+|S|)\right)\left(i+|S|\right)\delta _{i,S}$$
on $`\overline{M}_{g,n}`$.
Then, $`D|_{\overline{F}_{g,n}}`$ is trivial and $`D`$ has zero intersection with all strata of type (2.2.6). Moreover, $`D`$ has strictly positive intersection with all other strata if $`a>12b_{\mathrm{irr}}(g+n1)`$, $`2b_{\mathrm{irr}}>((n+g)/2)^2`$. If $`D`$ satisfies these inequalities, $`D`$ is nef. In particular $`F_1(\overline{M}_{g,n})`$ implies $`F_1(\overline{M}_{0,g+n}/S_g)`$.
Every nef line bundle on $`\overline{M}_{0,g+n}/S_g`$ is the pullback of a nef line bundle on $`\overline{M}_{g,n}`$. The strata numerically equivalent to a curve in the flag locus are precisely the strata of type (2.2.6).
###### Demonstration Proof
One checks by (2.1.6) that $`D`$ has trivial intersection with all the strata of type (2.2.6). These correspond to the strata of $`\overline{M}_{0,g+n}`$. Since the strata in $`\overline{M}_{0,n}`$ generate the Chow group, it follows that $`D`$ has trivial pullback, as in (0.3), to $`\overline{M}_{0,g+n}/S_g`$. Now one checks using (2.1-2.2) that $`D`$ has strictly positive intersection with all other strata if the inequalities in the statement hold. The final remark in the statement now follows.
Now suppose the inequalities of the statement hold. Then $`D`$ is nef by (0.3). Standard intersection formulae, e.g. \[Faber97\], show that
$$f^{}:\mathrm{Pic}(\overline{M}_{g,n})\mathrm{Pic}(\overline{M}_{0,g+n}/S_g)$$
is surjective. Suppose $`f^{}(W)`$ is nef. Then $`W+mD`$ will have non-negative intersection with all strata for $`m>>0`$ by the above. Thus it is nef by (0.3). โ
Of course the above theorem implies (0.7).
###### 4.8 Lemma
Let $`D=D_{g,n}`$ be the divisor of (4.7). Let $`h:\overline{M}_{j,s}\overline{M}_{g,n}`$ be a factor of one of the strata. If $`h`$ corresponds to a smooth rational component then $`h^{}(D)`$ is trivial. Otherwise $`๐ผ(h^{}(D))\overline{M}_{j,s}`$.
###### Demonstration Proof
In the first case we can assume (without affecting the pullback on Picard groups) that $`h`$ comes from a stratum in the flag locus, and so the result follows from (4.7).
Now assume either $`j1`$, or $`j=0`$ and the corresponding component contains at least one disconnecting node. If $`s1`$, let $`C`$ be the general fiber of one of the projections $`\overline{M}_{j,s}\overline{M}_{j,s1}`$. Then, $`C`$ is numerically equivalent to an effective sum of strata, not all of which are of type (2.2.6). Thus $`h^{}(D)C>0`$ by (4.7). Now (0.10) applies as long as $`j2`$.
Suppose $`j=1`$. Observe by (4.7) that $`dD_{j,s}h^{}(D)`$ is nef for sufficiently large $`d`$. Thus its enough to show $`๐ผ(D_{j,s})\overline{M}_{j,s}`$. By (4.7) and the proof of (3.3), $`D_{1,s}`$ is an effective sum of boundary divisors, with every divisor occurring with positive coefficient. Since the boundary supports an ample divisor, the result follows.
Now assume $`j=0`$, and the corresponding component contains a disconnecting node. As in the previous paragraph, we can assume $`h`$ defines $`\overline{R}_{g,n}`$, $`g1`$. By (4.7), for any divisor $`W`$, $`dDW`$ satisfies the hypothesis of (3.3). So, taking $`W`$ ample, the result follows from (4.1). โ
###### 4.9 Corollary
If $`D`$ is the nef divisor of (4.7), then
$$D^{}\overline{NE}_1(\overline{M}_{g,n})=\overline{NE}_1(\overline{M}_{0,g+n}/S_g).$$
###### Demonstration Proof
Let $`C\overline{M}_{g,n}`$ be a curve. By (4.8) and (1.2), $`DC=0`$ iff every moving component of the associated family of curves is smooth and rational, in which case $`C`$ is algebraically equivalent to a curve in $`\overline{F}_{g,n}`$. โ
## ยง5 Exceptional Loci
In this section we prove (0.9). We will use the following lemmas.
###### 5.1 Lemma
Let $`F`$ the fiber of $`\pi :\overline{M}_{g,n}\overline{M}_g`$ (resp. $`\pi :\overline{M}_{1,n}\overline{M}_{1,1}`$) over a point $`[C]`$ (resp. $`[C,p]`$) with $`C`$ smooth. Let $`G`$ be the automorphism group of $`C`$. Let $`D`$ be divisor on $`F`$ which is the restriction of a nef divisor in $`\mathrm{Pic}(\overline{M}_{g,n})`$. Let
$$f:FC^n/G$$
be the natural birational map. $`D`$ is big iff $`f_{}(D)`$ is big, in which case
$$๐ผ(D)f^1(๐ผ(f_{}(D)))F.$$
###### Demonstration Proof
We will abuse notation slightly and use the same symbol for a divisor on $`\overline{M}_{g,n}`$ and its restriction to $`F`$.
Suppose first $`g2`$. As in (3.4) we have an expression
$$D=\underset{iN}{}c_i\omega _i+\left(\underset{\frac{}{SNn2|S|2}}{}\left(\left(\underset{iS}{}c_i\right)b_{0,S}\right)\delta _{0,S}\right)$$
with $`c_i0`$, and with each $`\delta _{0,S}`$ having non-negative coefficient. Let
$$\mathrm{\Gamma }=\left(\underset{\frac{}{SNn2|S|2}}{}\left(\left(\underset{iS}{}c_i\right)b_{0,S}\right)\delta _{0,S}\right).$$
Observe that the exceptional divisors of $`f`$ are precisely the $`\delta _{0,S}`$ with $`|S|3`$, and that $`f(\delta _{0,S})f(\delta _{0,T})`$ iff $`TS`$. Thus the support of $`f^1(f_{}(\mathrm{\Gamma }))`$ is the union of the $`\delta _{0,S}`$ over subsets $`S`$ which contain a $`2`$-element subset, $`T`$, such that $`\delta _{0,T}`$ occurs in $`\mathrm{\Gamma }`$ with non-zero coefficient. By repeated application of (2.1.5), we obtain
$$(\underset{iS}{}c_i)b_{0,S}(\underset{iT}{}c_i)b_{0,T}$$
if $`TS`$. Thus the support of $`\mathrm{\Gamma }`$ contains the support of $`f^1(f_{}(\mathrm{\Gamma }))`$ (in fact since $`D`$ is $`f`$-nef, by negativity of contractions, the supports are the same). As each $`\omega _i`$ is a pullback along $`f`$ it follows that $`D`$ is big iff $`f_{}(D)`$ is big. By \[Keel99, 4.9\], $`\omega _i`$ is nef and $`๐ผ(\omega _i)F.`$ The result follows.
Now suppose $`g=1`$. By (3.3) we may express $`D`$ as
$$b_{0,S}\delta _{0,S}$$
with $`b_{0,S}0`$. Using (2.1) one checks $`b_{0,S}b_{0,T}`$ for $`TS`$. Now it follows as above that $`D`$ has the same support as $`f^1(f_{}(D))`$. The result follows. โ
###### 5.2 Lemma
Let $`C`$ be a smooth curve of genus $`g2`$, with automorphism group $`G`$. Let $`D`$ be an divisor on $`C^n/G`$ of the form
$$D=c_i\omega _i+\underset{ij}{}a_{ij}\mathrm{\Delta }_{ij}$$
with $`c_i,a_{ij}0`$ and $`c_i+c_ja_{ij}`$. $`D|_Z`$ is big for any subvariety $`ZC^n/G`$ not contained in any diagonal iff $`D`$ is $`q_i`$-relatively ample for each of the projections
$$q_i:C^n/GC^{n1}/G$$
given by dropping the $`i^{th}`$ point.
###### Demonstration Proof
The forward implication is obvious. Consider the reverse implication. As $`G`$ is finite itโs enough to consider the analogous statements on $`C^n`$. We assume $`D`$ is $`q_i`$-ample for all $`i`$. Write
$$D=c_i\omega _i+a_{ij}\mathrm{\Delta }_{ij}.$$
Reorder so that $`c_i=0`$ iff $`is+1`$. For each $`t>s`$, since $`D`$ is $`\pi _t`$-ample, $`a_{i(t)t}>0`$ for some $`i(t)t`$. Since $`c_i+c_ja_{ij}0`$, we must have $`i(t)s`$. If
$$p:C^nC^s$$
is the projection onto the first $`s`$ factors, then $`c_i\omega _i`$ is the pullback along $`p`$ of an ample divisor, while
$$\underset{t>s}{}a_{i(t)t}\mathrm{\Delta }_{i(t)t}$$
is $`p`$-ample. โ
###### 5.3 Lemma
Let $`C`$ be a smooth curve of genus $`1`$, with automorphism group $`G`$. Let
$$D=a_{ij}\mathrm{\Delta }_{ij}$$
be an effective combination of diagonals on $`C^n/G`$. Let $`HS_n`$ be the subgroup generated by the $`(ij)`$ with $`a_{ij}0`$. If $`H`$ is transitive, then $`D`$ is ample. If a partition $`N=SS^c`$ is preserved by $`H`$, then $`D`$ is pulled back along the fibration
$$C^n/GC^S/G\times C^{S^c}/G.$$
###### Demonstration Proof
The second claim is clear. For the first we induce on $`n`$. For $`n=2`$ the result is clear (the quotient is a curve and $`D`$ is non-trivial). Note since the genus is one, that all the diagonals (and hence $`D`$) are semi-ample, so its enough to show $`D`$ has positive intersection with every irreducible curve. But if $`\mathrm{\Delta }_{ij}E=0`$ for a curve $`E`$, then $`E`$ is numerically equivalent to a curve contained in $`\mathrm{\Delta }_{ij}`$ so itโs enough to show $`D|_{\mathrm{\Delta }_{ij}}`$ is ample, and for this we can apply induction. โ
###### Demonstration Proof of (0.9)
Suppose first $`g2`$. If $`n=0`$ and $`g=2`$, the result follows from \[Faber96\]. If $`n=0`$ and $`g3`$, we apply (3.5). If $`a=0`$, then $`D`$ is trivial by (2.1.1-2.1.4). Otherwise we have strict inequality in (3.5) by (2.1).
So we may assume $`n>0`$ and that $`D`$ is not pulled back from $`\overline{M}_{g,n1}`$. If $`D`$ is trivial on the general fiber of $`q:\overline{M}_{g,n}\overline{M}_{g,n1}`$, then it is $`q`$-trivial, since it is nef, thus, by an easy calculation, pulled back. So we can assume $`D`$ is $`q`$-big (for any choice of $`q`$).
Mimicking the argument of (3.3) we write $`D`$ as in (3.4)
$$D=\left(\underset{is}{}c_i\omega _i\right)+\left(\underset{\frac{}{S\{1,2,\mathrm{},n\}n2|S|2}}{}\left(\left(\underset{iS}{}c_i\right)b_{0,S}\right)\delta _{0,S}\right)+\pi ^{}(A)+E.$$
For any subvariety $`Z\overline{M}_g`$, $`A|_Z`$ is effective by (3.5). Thus by \[KollรกrMori98.3.23\] it is enough to show that $`๐ผ(D|_F)F\overline{M}_{g,n}`$ for $`F`$ as in (5.1). This is immediate from (5.1-5.2).
Now suppose $`g=1`$. Suppose first that $`D`$ is $`q_S`$ big for each fibration
$$q_S:\overline{M}_{1,n}\overline{M}_{1,S}\times _{\overline{M}_{1,1}}\overline{M}_{1,S^c}$$
and each non-trivial proper subset $`SN`$. We conclude as in the $`g2`$ case by applying (5.1) and (5.3).
So we may assume we have a non-trivial proper subset $`SN`$ such that $`D`$ has zero intersection with the general fiber (which is one dimensional) of $`q_S`$. As $`D`$ is nef, a short calculation shows $`D`$ is of the form $`\pi _S^{}(D_S)+\pi _{S^c}^{}(D_{S^c})`$ for divisors $`D_S`$ and $`D_{S^c}`$ on $`\overline{M}_{1,S}`$ and $`\overline{M}_{1,S^c}`$. Indeed one checks with (2.2.5) that
$$b_{0,TS}+b_{0,TS^c}=b_{0,T}$$
for all $`T`$, from which it follows that we may take
$`D_S`$ $`={\displaystyle \underset{TN}{}}b_{0,TS}\delta _{0,TS}`$
$`D_{S^c}`$ $`={\displaystyle \underset{TN}{}}b_{0,TS^c}\delta _{0,TS^c}.`$
Now note that $`\overline{R}_{1,m}`$ and all of the $`1`$-strata of $`\overline{M}_{1,n}`$ except for the elliptic tails lie over the point $`\delta _{\mathrm{irr}}\overline{M}_{1,1}`$. It follows from (3.6) that for divisors $`D_S\mathrm{Pic}(\overline{M}_{1,S})`$ and $`D_{S^c}\mathrm{Pic}(\overline{M}_{1,S^c})`$, if $`\pi _S^{}D_S+\pi _{S^c}^{}D_{S^c}`$ is nef, then both $`D_S`$ and $`D_{S^c}`$ are nef, as long as each has non-negative intersection with a family of elliptic tails. Let $`E\overline{M}_{1,n}`$ be a family of elliptic curves. Then so are $`E_S=\pi _S(E)`$ and $`E_{S^c}=\pi _{S^c}(E)`$ and $`\delta _{\mathrm{irr}}E_S=\delta _{\mathrm{irr}}E_{S^c}>0`$. So if we define $`t`$ by
$$(D_S+t\delta _{\mathrm{irr}})E_S=0$$
then
$$(D_{S^c}t\delta _{\mathrm{irr}})E_{S^c}=D_SE_S+D_{S^c}E_{S^c}=DE0.$$
Hence
$$D=\pi _S^{}(D_S+t\delta _{\mathrm{irr}})+\pi _{S^c}^{}(D_{S^c}t\delta _{\mathrm{irr}})$$
with both $`D_S+t\delta _{\mathrm{irr}}`$ and $`D_{S^c}t\delta _{\mathrm{irr}}`$ nef. โ
###### Remark Remarks
A natural analog of $`f`$ from (5.1) for $`g=0`$ is the birational map
$$f:\overline{M}_{0,n}Q$$
with $`Q`$ one of the maximal GIT quotients of $`(^1)^{\times n}`$ (e.g. the unique $`S_n`$ symmetric quotient). The birational geometry of $`Q`$ is well understood, see \[HuKeel00\], and in particular the analog of (5.3) holds. Thus if one could prove the analog of (5.1), one could extend (0.9) to all genera. This would imply (0.2).
## ยง6 Ad hoc examples
###### 6.1 Proposition
(char $`0`$) Let $`D`$ be an $`F`$-nef divisor $`a\lambda b_i\delta _i`$ on $`\overline{M}_g`$. Assume further that for each coefficient $`b_i`$, $`[g/2]i1`$ that either $`b_i=0`$ or $`b_ib_{\mathrm{irr}}`$. Then $`D`$ is nef.
###### Demonstration Proof
By (0.3) it is enough to show $`h^{}(D)`$ is nef, for the composition $`h:\overline{M}_{0,2g}\overline{R}_{g,n}\overline{M}_g`$. We can assume $`b_{\mathrm{irr}}>0`$ (otherwise all $`b_i=0`$ ). Suppose first $`b_i=0`$ for some $`i>0`$. Consider the map
$$\overline{M}_{i,1}\overline{M}_g$$
(from the usual product decomposition of $`\delta _i`$). Let $`D^{}`$ be the pullback of $`D`$, which we express in the usual basis. Intersecting with an appropriate stratum of type (2.1.5) on $`\overline{M}_{i,1}`$ gives the inequality
$$c+b_{j,\mathrm{}}b_{j,\{1\}}=b_{gj,\mathrm{}},g>j>0$$
(where $`c=c_1`$). In our case $`c=b_i=0`$ and so $`b_{j,\mathrm{}}=b_{j,\{1\}}`$, and thus $`D^{}`$ is pulled back from $`\overline{M}_{i,0}`$. The same argument applies to the other factor of $`\delta _i`$. Thus by induction on $`g`$ we may assume $`D|_{\delta _i}`$ is nef for any $`i>0`$ with $`b_i=0`$.
Using $`12\lambda =\kappa +\delta `$ and the expression $`K_{\overline{M}_{0,n}}=\kappa +\delta `$ we have that
$$f^{}(D/b_{\mathrm{irr}})=K+\underset{i=1}{\overset{[g/2]}{}}(2b_i/b_{\mathrm{irr}})\delta _i+B$$
where $`\delta _i`$ on $`\overline{M}_{0,2g}`$ indicates the pullback of $`\delta _i`$ on $`\overline{M}_g`$ (note the pullback of $`\delta _i`$ to $`\overline{R}_{g,n}`$ is effective and irreducible) which is an effective sum of boundary divisors with coefficient one (or zero), and $`B`$ is the sum of the boundary divisors of $`\overline{M}_{0,2g}`$ other than the $`\delta _i`$. Let $`\mathrm{\Delta }=_{i=1}^{[g/2]}(2b_i/b_{\mathrm{irr}})\delta _i+B`$. By (2.1.4) the coefficients of $`\mathrm{\Delta }`$ are non-negative and, by the above, $`K+\mathrm{\Delta }`$ is nef on any boundary component whose coefficient in $`\mathrm{\Delta }`$ is greater than one. Suppose $`D`$ (or equivalently $`K+\mathrm{\Delta }`$) is not nef. Then there is a $`K+\mathrm{\Delta }`$ negative extremal ray $`R`$ of $`\overline{NE}_1(\overline{M}_{0,2g})`$. If $`\mathrm{\Delta }^{}`$ is obtained from $`\mathrm{\Delta }`$ by shrinking to one the coefficients that are greater than one, then by the above $`(K+\mathrm{\Delta }^{})R<0`$. Thus by \[KeelMcKernan96, 1.2\], $`R`$ is generated by a one dimensional stratum, a contradiction. โ
Note (6.1) applies to all the edges of the Faber cone computed in \[Faber96\] for $`g7`$, giving another proof of (0.2) for $`n=0`$, $`g7`$. Unfortunately, this fails for $`g=10`$ where $`30\lambda 3\delta _06\delta _16\delta _22\delta _34\delta _46\delta _5`$ is a vertex. It also answers in the affirmative question (a) at the end of ยง2 of \[Faber96\].
###### 6.2 Corollary
(char $`0`$) The ray $`10\lambda 2\delta +\delta _{\mathrm{irr}}`$ is nef on $`\overline{M}_g`$ for all $`g2`$.
Also, by averaging with the nef ray $`12\lambda \delta _{\mathrm{irr}}`$, it recovers the ampleness result of \[Cornalba-Harris88\].
###### 6.3 Corollary (Cornalba-Harris)
(char $`0`$) The class $`11\lambda \delta `$ is nef on $`\overline{M}_g`$ for all $`g2`$.
The ray $`11\lambda \delta `$ is known to be semi-ample and the associated map to be a birational contraction of relative Picard number one with exceptional locus $`\mathrm{\Delta }_1`$ which blows down families of elliptic tails. In fact, except for some special low genus examples, this is the only elementary divisorial blowdown of $`\overline{M}_g`$:
###### 6.4 Proposition
(char $`0`$, $`g5`$) The only divisorial contraction $`f:\overline{M}_{g,n}X`$ of relative Picard number one with $`X`$ projective is the blowdown of the elliptic tails.
###### Demonstration Proof
By (0.9) some boundary divisor is $`f`$-exceptional. We assume $`\delta _i`$ is exceptional, for some $`i1`$, the argument for $`\delta _{\mathrm{irr}}`$ is only notationally different. By the product decomposition, $`f`$ induces a fibration on $`\overline{M}_{k,1}`$ either for $`k=i`$ or $`k=gi`$. Assume $`f`$ does not contract the elliptic tails. Then by (0.9), this induced fibration factors through the map $`\pi _k:\overline{M}_{k,1}\overline{M}_k`$. For $`i>1`$, the general fibers of $`\pi _i`$ and $`\pi _{gi}`$ generate the same ray in $`N_1(\overline{M}_g)`$ (indeed, the only basic class with non-zero intersection with either is $`\delta _i`$, which intersects either negatively). Thus for $`i>1`$, $`f`$ induces a fibration on either factor, and contracts any curve contracted by $`\pi _k`$, $`k=i,gi`$. But for $`k>2`$ (which holds for at least one of $`i`$ or $`gi`$ since $`g5`$) the image in $`N_1(\overline{M}_g)`$ of $`N_1(\pi _k)`$ (the relative Neron-Severi group of curves of $`\pi _k`$ ) has dimension at least two. Indeed the fiber of $`\pi _k`$ over a general point of $`\delta _1(\overline{M}_k)`$ is a curve with two irreducible components, and a quick computation shows the images in $`N_1(\overline{M}_g)`$ of these components are linearly independent ($`\delta _1`$ intersects one negatively, and the other positively). โ
###### Remark Remark
Bill Rulla, \[Rulla00\] has shown that on $`\overline{M}_3`$ there is a second relative Picard number one contraction of $`\delta _1`$. The corresponding extremal ray is generated by the curve (2.2.4), with $`i=1`$.
## ยง7 Reformulation of $`F_1(\overline{M}_{0,n})`$ and review of the evidence
Let $`S\overline{M}_{0,n}`$ be the union, with reduced structure, of the two dimensional strata. The irreducible components of $`S`$ are all smooth Del Pezzo surfaces, either $`\overline{M}_{0,5}`$ or $`^1\times ^1`$.
###### 7.1 Proposition
The natural map $`A_1(S)A_1(\overline{M}_{0,n})`$ is an isomorphism, and under this map $`\overline{NE}_1(S)`$ is identified with the Faber cone. Thus $`F_1(\overline{M}_{0,n})`$ is equivalent to the statement that the natural injection $`\overline{NE}_1(S)\overline{NE}_1(\overline{M}_{0,n})`$ is an isomorphism.
###### Demonstration Proof
The strata generate the Mori cone of $`S`$, so itโs enough to prove the statement about Chow groups. This follows from the Chow group description given in \[KontsevichManin96\]. โ
It is known that each stratum of $`\overline{M}_{0,n}`$ generates an extremal ray. Each can be contracted by a birational morphism of relative Picard number one which is in fact a log Mori fibration( cf. \[KeelMckernan96\]). The main evidence for $`F_1(\overline{M}_{0,n})`$ is \[KeelMckernan96, 1.2\], which states in particular that (0.2) holds for any extremal ray of $`\overline{NE}_1(\overline{M}_{0,n})`$ that can be contracted by a morphism of relative Picard number one, except possibly when the map is birational and has one dimensional exceptional locus. Any stratum deforms once $`n9`$, so (0.2) would imply this extra condition always holds for such $`n`$. In practice the main way of showing a curve generates an extremal way is to exhibit a corresponding contraction, so if (0.2) fails, it will be a challenge to find a counter-example.
Kapranov proved that $`\overline{M}_{0,n}`$ is the inverse limit of all GIT quotients $`Q`$ (varying the polarization) for $`\mathrm{SL}_2`$ acting diagonally on $`(^1)^n`$. In particular each $`Q`$ is a contraction of $`\overline{M}_{0,n}`$, and so inherits a stratification. In \[HuKeel00\] $`F_1(Q)`$ and $`F^1(Q)`$ are proved. One can also consider the contractions $`\overline{M}_{0,n}\overline{L}_{n2}`$, introduced in \[LosevManin00\]. $`\overline{L}_{n2}`$ is a smooth projective toric variety. On any projective toric variety the toric strata generate the cones of effective cycles. The toric strata for $`\overline{L}_{n2}`$ are images of strata in $`\overline{M}_{0,n}`$, thus $`F_k(\overline{L}_{n2})`$ holds for all $`k`$. We remark that $`\overline{L}_{n2}`$ gives particularly good evidence for $`F_1(\overline{M}_{0,n})`$ as it has Picard number $`2^{n2}n+1`$, roughly half the Picard number of $`\overline{M}_{0,n}`$ whereas a GIT quotient $`Q`$ can have Picard number at most $`n`$.
|
warning/0006/quant-ph0006121.html
|
ar5iv
|
text
|
# Contents
## 1 Introduction
As is already known from classical optics, the use of instruments in optical experiments needs careful examination with regard to their action on the light under study. Fibres, beam splitters, cavities, spectral filters etc. are familiar examples of optical instruments, which are typically built up by dielectric bodies. In quantum optics an important aspect is the influence of such material bodies on the quantum features of light including its interaction with (microscopic) atomic systems. For example, let us consider a $`50\%/50\%`$ beam splitter oriented at $`45^{}`$ to an incident light beam. In classical optics the beam splitter simply divides the incoming beam into two (apart from a phase shift) equal outgoing parts propagating perpendicular to each other, and with the same scaling factor the classical noise of the incident field is transferred to the two fields in the output channels of the beam splitter. It is intuitively clear that in quantum optics the noise of the vacuum in the unused input port of the beam splitter introduces additional noise in the two output beams and thus the quantum statistics of the output field may differ significantly from that of the input field provided that the input field is prepared in a nonclassical state.
Moreover, when two light beams that are prepared in non-classical states are superimposed by the beam splitter, then the outgoing field is prepared in a nonclassically correlated bipartite state, also called an entangled state. Entanglement as a typical quantum coherence feature plays an important role in quantum communication. Let us assume that two light beams have been prepared in an entangled state. In order to use them, e.g., in quantum teleportation, the beams should be transmitted through optical channels such as fibres. Here the crucial point is to what extent the entanglement can be preserved during the propagation of the beams, because in any real fibre the absorption gives necessarily rise to an entanglement degradation.
From a more theoretical point of view, a very interesting question is that of the Casimir force between material bodies. A crude physical explanation of the Casimir force is that the vacuum energies of two regions spatially separated by the bodies differ due to the presence of the matter, which in a rough approximation can be described simply by a boundary condition on the quantization volume. In fact, this simple model does not take account of the dispersive and absorptive properties of the matter and fails in the high-frequency limit where the matter becomes transparent.
These few examples show that it is necessary to include in the theory the presence of material bodies when considering the quantized radiation field and its interaction with atoms. In principle, material bodies could be included as a part of the matter to which the radiation field is coupled and treated microscopically. However, there is a class of material bodies whose action can be included in the quantum theory exactly, namely dielectric bodies that respond linearly to the electromagnetic field, the response being described in terms of a phenomenologically introduced space-dependent dielectric permittivity. Such a concept has โ similar to classical optics โ the benefit of being universally valid, because it uses only general physical properties, without the need of involved ab initio calculations.
The quantum theory of radiation in the presence of dielectric matter has been studied over a long period. Quantization of the electromagnetic field in dielectric media with assumed real and frequency-independent permittivity has been treated extensively . In the same context, dispersive dielectrics have been considered and attempts have been made to extend the concepts to nonlinear media . However, it is well known that the permittivity is a complex function of frequency which has to satisfy the KramersโKronig relations, which state that the real part of the permittivity (responsible for dispersion) and the imaginary part (responsible for absorption) are necessarily connected with each other. A consequence of the existence of the imaginary part of the permittivity is that the commonly used mode expansion of the (macroscopic) electromagnetic field fails, at least in frequency intervals where the absorption cannot be disregarded. Obviously, an expansion of the field in terms of damped (non-orthogonal) waves would not be complete. From statistical mechanics it is clear that dissipation is unavoidably connected with the appearance of a random force which gives rise to an additional noise source of the electromagnetic field. Hence, any quantum theory that is based on the assumption of a real permittivity can only be valid for narrow-bandwidth fields far from medium resonances where absorption can safely be disregarded.
For the last years there has been an increasing number of articles dealing with the problem of the formulation of quantum electrodynamics in dielectric media of given complex permittivity satisfying the KramersโKronig relations . A systematic and quantum-theoretically consistent approach to the problem of the quantization of the radiation field in absorbing bulk dielectrics is given in on the basis of the microscopic Hopfield model of a dielectric . It is based on an explicit Fano-type diagonalization of a Hamiltonian consisting of the electromagnetic field, a (harmonic-oscillator) polarization field representing the dielectric matter, and a continuous set of (harmonic-oscillator) reservoir variables accounting for absorption. The resulting expression for the vector potential can be written in terms of the Green tensor of the classical scattering problem, which, in fact, makes it possible to perform the quantization of the electromagnetic field in the presence of arbitrary dielectric bodies of (phenomenologically) given permittivities , without referring to specific microscopic models of the bodies, which were hard to establish for general systems. The concept is based on a source-quantity representation of the electromagnetic field, in which the electromagnetic-field operators are expressed in terms of a continuous set of fundamental bosonic fields via the Green tensor of the classical problem.
Let us give a brief guide to the topics covered. After giving an outline of the quantization scheme based on the microscopic Hopfield model of a dielectric bulk material (Sec. 2), we show how the classical phenomenological Maxwell equations of the electromagnetic field in the presence of dielectric matter of given space- and frequency-dependent permittivity can be transferred to quantum theory (Sec. 3). For this purpose we first summarize the basic properties of the classical Maxwell equations and express the electromagnetic field in terms of the Green tensor and a continuous set of appropriately chosen dynamical field variables (Sec. 3.1). We then perform the quantization by identifying the dynamical field variables with bosonic fields associated with the elementary excitations of the composed system (Sec. 3.2). Having quantized the electromagnetic field, the question arises of how to include in the theory the interaction of the medium-assisted field with atomic systems (Sec. 4). In order to answer it, we present both the minimal-coupling Hamiltonian (Sec. 4.1) and the multipolar-coupling Hamiltonian (Sec. 4.2) in the Coulomb gauge. To illustrate the basic theoretical concept, we discuss a number of applications such as the inputโoutput relations of radiation (Sec. 5.1) and the transformation of radiation-field quantum states (Sec. 5.2) at absorbing four-port devices, and the spontaneous decay of an excited atom near the (planar) surface of an absorbing body (Sec. 6.2) and in a (spherical) micro-cavity with intrinsic material losses (Secs. 6.3 and 6.4). For the sake of transparency, we perform the calculations for isotropic media and outline the extension to other media in a separate section at the end (Sec. 7).
## 2 Hopfield model and Fano diagonalization
Following the quantization scheme in , we consider a Hopfield model of a bulk dielectric in which $`N`$ harmonic oscillator fields describing the polarization of the dielectric medium are linearly coupled to a continuum of harmonic oscillators standing for a reservoir. Such a model leads to an energy flow essentially only in one direction, namely from the medium to the reservoir where it โdisappearsโ, hence it is absorbed. The overall system that consists of the radiation, the dielectric-medium polarization, the reservoir, and couplings between them may be regarded as being a Hamiltonian system whose Lagrangian reads
$$L=\mathrm{d}^3๐ซ=\mathrm{d}^3๐ซ(_{\mathrm{rad}}+_{\mathrm{mat}}+_{\mathrm{int}}),$$
(2.1)
where
$$_{\mathrm{rad}}=\frac{1}{2}\epsilon _0\left[(\dot{๐}+\mathbf{}U)^2c^2(\mathbf{}\times ๐)^2\right]$$
(2.2)
($`U`$, scalar potential; $`๐`$, vector potential; $`c^2`$ $`=`$ $`\epsilon _0\mu _0`$), and
$$_{\mathrm{mat}}=\underset{i=1}{\overset{N}{}}\frac{1}{2}\mu \left(\dot{๐}_i^2\omega _i^2๐_i^2\right)+\underset{0}{\overset{\mathrm{}}{}}d\omega \frac{1}{2}\mu \left[\dot{๐}^2(\omega )\omega ^2๐^2(\omega )\right],$$
(2.3)
$$_{\mathrm{int}}=\underset{i=1}{\overset{N}{}}\alpha _i\left(๐\dot{๐}_i+U\mathbf{}๐_i\right)\underset{0}{\overset{\mathrm{}}{}}d\omega \dot{๐}(\omega )\underset{i=1}{\overset{N}{}}v_i(\omega )๐_i.$$
(2.4)
Here $`_{\mathrm{rad}}`$ and $`_{\mathrm{mat}}`$ are respectively the free Lagrangian densities of the radiation field<sup>1</sup><sup>1</sup>1Note that the vector potential fulfills the Coulomb-gauge condition $`\mathbf{}๐`$ $`=`$ $`0`$. and the matter fields \[i.e., the medium oscillator fields $`๐_i`$ and the reservoir oscillator fields $`๐(\omega )`$ with density $`\mu `$\], and $`_{\mathrm{int}}`$ is the interacting part, where the mediumโfield coupling constants $`\alpha _i`$ play the role of (electric) polarizabilities, and the mediumโreservoir coupling constants $`v_i(\omega )`$ are assumed to be square-integrable functions of $`\omega `$. Introducing the canonical momenta
$$๐ท=\frac{}{\dot{๐}}=\epsilon _0\dot{๐},$$
(2.5)
$$๐_i=\frac{}{\dot{๐}_i}=\mu \dot{๐}_i\alpha _i๐,$$
(2.6)
$$๐(\omega )=\frac{}{\dot{๐}(\omega )}=\mu \dot{๐}(\omega )\underset{i=1}{\overset{N}{}}v_i(\omega )๐_i,$$
(2.7)
it is not difficult to perform the Legendre transformation and construct the Hamiltonian $`H`$ $`=`$ $`H_{\mathrm{rad}}`$ $`+`$ $`H_{\mathrm{mat}}`$ $`+`$ $`H_{\mathrm{int}}`$ of the overall system.
Since the dielectric medium is assumed to be infinitely extended, it is convenient to go to the reciprocal space<sup>2</sup><sup>2</sup>2The spatial Fourier transform $`\stackrel{~}{F}(๐ค)`$ of a function $`F(๐ซ)`$ is defined according to the relation $`F(๐ซ)`$ $`=`$ $`(2\pi )^{3/2}\mathrm{d}^3๐ค\stackrel{~}{F}(๐ค)\mathrm{e}^{i\mathrm{๐ค๐ซ}}`$.,
$$๐(๐ซ)\stackrel{~}{๐}(๐ค)=\underset{\lambda =1}{\overset{2}{}}A_\lambda (๐ค)๐_\lambda (๐ค),$$
(2.8)
$$๐_i(๐ซ)\stackrel{~}{๐}_i(๐ค)=X_i(๐ค)๐ฟ+\underset{\lambda =1}{\overset{2}{}}X_{i\lambda }(๐ค)๐_\lambda (๐ค)$$
(2.9)
\[$`๐ฟ`$ $`=`$ $`๐ค/|๐ค|`$, $`๐_\lambda (๐ค)๐ค`$\], and $`\stackrel{~}{๐ท}(๐ค)`$, $`\stackrel{~}{๐}_i(๐ค)`$, $`\stackrel{~}{๐}(๐ค,\omega )`$, $`\stackrel{~}{๐}(๐ค,\omega )`$ accordingly, and to introduce, with regard to quantization, the new variables
$`a_\lambda (๐ค)`$ $`=`$ $`\sqrt{{\displaystyle \frac{\epsilon _0}{2\mathrm{}\stackrel{~}{k}c}}}\left[\stackrel{~}{k}cA_\lambda (๐ค)+{\displaystyle \frac{i}{\epsilon _0}}\mathrm{\Pi }_\lambda (๐ค)\right],`$ (2.10)
$`b_{i\lambda }(๐ค)`$ $`=`$ $`\sqrt{{\displaystyle \frac{\mu }{2\mathrm{}\stackrel{~}{\omega _i}}}}\left[\stackrel{~}{\omega }_iX_{i\lambda }(๐ค)+{\displaystyle \frac{i}{\mu }}Q_{i\lambda }(๐ค)\right],`$ (2.11)
$`b_\lambda (๐ค,\omega )`$ $`=`$ $`\sqrt{{\displaystyle \frac{\mu }{2\mathrm{}\omega }}}\left[i\omega X_\lambda (๐ค,\omega )+{\displaystyle \frac{1}{\mu }}Q_\lambda (๐ค,\omega )\right],`$ (2.12)
with
$$\stackrel{~}{k}^2=k^2+\underset{i=1}{\overset{N}{}}\frac{\alpha _i^2}{\mu \epsilon _0c^2},\stackrel{~}{\omega _i}^2=\omega _i^2+_0^{\mathrm{}}d\omega \frac{v_i^2(\omega )}{\mu ^2},$$
(2.13)
and
$`b_i(๐ค)`$ $`=`$ $`\sqrt{{\displaystyle \frac{\mu }{2\mathrm{}\stackrel{~}{\omega }_i}}}\left[\stackrel{~}{\omega }_iX_i(๐ค)+{\displaystyle \frac{i}{\mu }}Q_i(๐ค)\right],`$ (2.14)
$`b_{}(๐ค,\omega )`$ $`=`$ $`\sqrt{{\displaystyle \frac{\mu }{2\mathrm{}\omega }}}\left[i\omega X_{}(๐ค,\omega )+{\displaystyle \frac{1}{\mu }}Q_{}(๐ค,\omega )\right]`$ (2.15)
\[$`\stackrel{~}{\omega }_i^2`$ $`=`$ $`\stackrel{~}{\omega }_i^2+\alpha _i^2/(\mu ฯต_0)`$\].
The fields are now quantized in the familiar way by conversion of the complex amplitudes $`a_\lambda (๐ค)`$, $`b_{i\lambda ()}(๐ค)`$, and $`b_{\lambda ()}(๐ค,\omega )`$ into bosonic annihilation operators $`\widehat{a}_\lambda (๐ค)`$, $`\widehat{b}_{i\lambda ()}(๐ค)`$, and $`\widehat{b}_{\lambda ()}(๐ค,\omega )`$, and the Hamiltonian can be expressed in terms of the annihilation and creation operators as follows:
$$\widehat{H}=\widehat{H}^{}+\widehat{H}_{\mathrm{mat}}^{},\widehat{H}^{}=\widehat{H}_{\mathrm{rad}}+\widehat{H}_{\mathrm{mat}}^{}+\widehat{H}_{\mathrm{int}}^{},$$
(2.16)
$`\widehat{H}_{\mathrm{mat}}^{}={\displaystyle }\mathrm{d}^3๐ค\{{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}\mathrm{d}\omega \mathrm{}\omega \widehat{b}_{}^{}(๐ค,\omega )\widehat{b}_{}(๐ค,\omega )+{\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{}\stackrel{~}{\omega }_i\widehat{b}_i^{}(๐ค)\widehat{b}_i(๐ค)`$ (2.17)
$`+{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}\mathrm{d}\omega {\displaystyle \underset{i=1}{\overset{N}{}}}\left[\frac{1}{2}\mathrm{}V_i(\omega )(\widehat{b}_i^{}(๐ค)+\widehat{b}_i(๐ค))(\widehat{b}_{}^{}(๐ค,\omega )+\widehat{b}_{}(๐ค,\omega ))\right]\},`$
$$\widehat{H}_{\mathrm{rad}}^{}=\underset{\lambda =1}{\overset{2}{}}\mathrm{d}^3๐ค\mathrm{}\stackrel{~}{k}c\widehat{a}_\lambda ^{}(๐ค)\widehat{a}_\lambda (๐ค),$$
(2.18)
$`\widehat{H}_{\mathrm{mat}}^{}={\displaystyle \underset{\lambda =1}{\overset{2}{}}}{\displaystyle }\mathrm{d}^3๐ค\{{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}\mathrm{d}\omega \mathrm{}\omega \widehat{b}_\lambda ^{}(๐ค,\omega )\widehat{b}_\lambda (๐ค,\omega )+{\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{}\stackrel{~}{\omega }_i\widehat{b}_{i\lambda }^{}(๐ค)\widehat{b}_{i\lambda }(๐ค)`$ (2.19)
$`+{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}\mathrm{d}\omega {\displaystyle \underset{i=1}{\overset{N}{}}}[\frac{1}{2}\mathrm{}V_i(\omega )(\widehat{b}_{i\lambda }^{}(๐ค)+\widehat{b}_{i\lambda }(๐ค))(\widehat{b}_\lambda ^{}(๐ค,\omega )+\widehat{b}_\lambda (๐ค,\omega ))`$
$`+{\displaystyle \underset{ji}{}}\frac{1}{2}\mathrm{}V_i(\omega )V_j(\omega )(\widehat{b}_{i\lambda }^{}(๐ค)+\widehat{b}_{i\lambda }(๐ค))(\widehat{b}_{j\lambda }^{}(๐ค)+\widehat{b}_{j\lambda }(๐ค))]\},`$
$$\widehat{H}_{\mathrm{int}}^{}=\underset{\lambda =1}{\overset{2}{}}\underset{i=1}{\overset{N}{}}\mathrm{d}^3๐ค\frac{1}{2}i\mathrm{}\mathrm{\Lambda }_i(k)\left[\widehat{a}_\lambda ^{}(๐ค)+\widehat{a}_\lambda (๐ค)\right]\left[\widehat{b}_{i\lambda }^{}(๐ค)+\widehat{b}_{i\lambda }(๐ค)\right],$$
(2.20)
where $`V_i(\omega )`$ $`=`$ $`(v_i(\omega )/\mu )(\omega /\stackrel{~}{\omega }_i)^{1/2}`$ and $`\mathrm{\Lambda }_i(k)`$ $`=`$ $`[(\stackrel{~}{\omega }_i\alpha _i^2)/(\mu c\epsilon _0\stackrel{~}{k})]^{1/2}`$. The bilinear Hamiltonian can be diagonalized (separately for the transverse and longitudinal parts) by applying a Fano-type technique such that<sup>3</sup><sup>3</sup>3A different technique for diagonalization is utilized in , where path-integral quantization of the Lagrangian (2.1) is performed.
$$\widehat{H}_{\mathrm{mat}}^{}=\mathrm{d}^3๐ค\underset{0}{\overset{\mathrm{}}{}}d\omega \mathrm{}\omega \widehat{B}_{}^{}(๐ค,\omega )\widehat{B}_{}(๐ค,\omega ),$$
(2.21)
$$\widehat{H}^{}=\underset{\lambda =1}{\overset{2}{}}\mathrm{d}^3๐ค\underset{0}{\overset{\mathrm{}}{}}d\omega \mathrm{}\omega \widehat{C}_\lambda ^{}(๐ค,\omega )\widehat{C}_\lambda (๐ค,\omega ).$$
(2.22)
Since the derivation of the formulas for expressing the old bosonic operators $`\widehat{b}_i(๐ค)`$, $`\widehat{b}_i^{}(๐ค)`$, $`\widehat{b}_{}(๐ค,\omega )`$, $`\widehat{b}_{}^{}(๐ค,\omega )`$ and $`\widehat{a}_\lambda (๐ค,\omega )`$, $`\widehat{a}_\lambda ^{}(๐ค,\omega )`$, $`\widehat{b}_{i\lambda }(๐ค)`$, $`\widehat{b}_{i\lambda }^{}(๐ค)`$, $`\widehat{b}_\lambda (๐ค,\omega )`$, $`\widehat{b}_\lambda ^{}(๐ค,\omega )`$ in terms of the new bosonic operators $`\widehat{B}_{}(๐ค,\omega )`$, $`\widehat{B}_{}^{}(๐ค,\omega )`$ and $`\widehat{C}_\lambda (๐ค,\omega )`$, $`\widehat{C}_\lambda ^{}(๐ค,\omega )`$, respectively, is rather lengthy, we renounce it here and refer the reader to .
The vector potential and the transverse part of the medium polarization can then be expressed in terms of the (polariton-like) operators $`\widehat{C}_\lambda (๐ค,\omega )`$, $`\widehat{C}_\lambda ^{}(๐ค,\omega )`$ as
$$\widehat{\stackrel{~}{๐}}(๐ค)=\underset{\lambda =1}{\overset{2}{}}๐_\lambda (๐ค)\sqrt{\frac{\mathrm{}}{\pi \epsilon _0}}\underset{0}{\overset{\mathrm{}}{}}d\omega \omega \sqrt{\epsilon _\mathrm{I}(\omega )}\stackrel{~}{G}(๐ค,\omega )\widehat{C}_\lambda (๐ค,\omega )+\text{H.c.}$$
(2.23)
and
$`\widehat{\stackrel{~}{๐}}(๐ค)={\displaystyle \underset{\lambda =1}{\overset{2}{}}}{\displaystyle \underset{i=1}{\overset{N}{}}}\alpha _i\widehat{๐}_{i\lambda }(๐ค)=i{\displaystyle \underset{\lambda =1}{\overset{2}{}}}๐_\lambda (๐ค)\sqrt{{\displaystyle \frac{\mathrm{}\epsilon _0}{\pi }}}`$ (2.24)
$`\times [{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}\mathrm{d}\omega \omega ^2[\epsilon (\omega )1]\sqrt{\epsilon _\mathrm{I}(\omega )}\stackrel{~}{G}(๐ค,\omega )\widehat{C}_\lambda (๐ค,\omega )+\mathrm{H}.\mathrm{c}.]+\widehat{\stackrel{~}{๐}}_\mathrm{N}(๐ค),`$
$$\widehat{\stackrel{~}{๐}}_\mathrm{N}(๐ค)=\underset{\lambda =1}{\overset{2}{}}๐_\lambda (๐ค)\sqrt{\frac{\mathrm{}\epsilon _0}{\pi }}\underset{0}{\overset{\mathrm{}}{}}d\omega i\sqrt{\epsilon _\mathrm{I}(\omega )}\widehat{C}_\lambda (๐ค,\omega )+\mathrm{H}.\mathrm{c}..$$
(2.25)
Here, $`\epsilon (\omega )`$ $`=`$ $`\epsilon _\mathrm{R}(\omega )`$ $`+`$ $`i\epsilon _\mathrm{I}(\omega )`$ is the complex (model) permittivity of the medium, and
$$\stackrel{~}{G}(๐ค,\omega )=\frac{c^2}{\omega ^2\epsilon (\omega )k^2c^2}$$
(2.26)
is the Green function of the classical Maxwell equations with that permittivity. From Eq. (2.24) \[together with Eq. (2.25)\] it is seen that the (transverse) polarization of the medium consists of two qualitatively different terms. Obviously, the first term is the induced polarization, whose frequency components are given by the frequency components of the electric field strength of the radiation multiplied by $`\epsilon _0[\epsilon (\omega )`$ $``$ $`1]`$, and the second term is the (noise) polarization, i.e., the fluctuating component of the polarization that is associated with absorption.
Recalling that the longitudinal part of the electric field operator is given by $`ฯต_0\widehat{\stackrel{~}{๐}}_{}(๐ค)`$ $`=`$ $`_i\alpha _i\widehat{\stackrel{~}{X}}_i(๐ค)๐ฟ`$ and recalling the transversality of the displacement operator $`\widehat{\stackrel{~}{๐}}(๐ค)`$, a relation between $`\widehat{\stackrel{~}{๐}}_{}`$ and a longitudinal (noise) polarization defined analogously to Eq. (2.25) \[by replacing $`\widehat{C}_\lambda (๐ค,\omega )`$ with $`\widehat{B}_{}(๐ค,\omega )`$\] can be derived.
In summary, the quantized electromagnetic field can be expressed, via a source-quantity representation with the classical Green function, in terms of the permittivity and a continuum of harmonic oscillators. It is worth noting that in this formulation there is no explicit hint at the underlying microscopic model. Thus, it seems quite natural to generalize the theory to arbitrary dielectric matter of given space- and frequency-dependent permittivity by transferring the classical source-quantity representation of the electromagnetic field directly to quantum theory.
## 3 The medium-assisted Maxwell field
From now we will not refer to one or the other microscopic model of the dielectric media. Instead we will start from the familiar phenomenological Maxwell equations, assuming that the permittivity is known, e.g., from measurements. In order to allow for an arbitrary formation of different (non-moving) dielectric bodies in space, we will assume that the permittivity varies with space. For the sake of transparency we will disregard the tensor character of the permittivity, restricting our attention to isotropic media (for the extension to anisotropic media, see Sec. 7.2).
### 3.1 Classical basic equations
Let us first briefly outline the classical theory and bring it in a form suitable for quantization. The phenomenological Maxwell equations of the electromagnetic field in the presence of dielectric bodies but without additional charge and current densities read
$$\mathbf{}๐(๐ซ)=0,$$
(3.1)
$$\mathbf{}\times ๐(๐ซ)+\dot{๐}(๐ซ)=0,$$
(3.2)
$$\mathbf{}๐(๐ซ)=0,$$
(3.3)
$$\mathbf{}\times ๐(๐ซ)\dot{๐}(๐ซ)=0,$$
(3.4)
where the displacement field $`๐`$ is related to the electric field $`๐`$ and the polarization field $`๐`$ according to
$$๐(๐ซ)=\epsilon _0๐(๐ซ)+๐(๐ซ),$$
(3.5)
and for nonmagnetic matter it may be assumed that
$$๐(๐ซ)=\frac{1}{\mu _0}๐(๐ซ)$$
(3.6)
(for the extension to magnetic matter, see Sec. 7.3).
Let us consider arbitrarily inhomogeneous (isotropic) media and assume that the polarization responds linearly and locally to the electric field. In this case, the most general relation between the polarization and the electric field which is in agreement with the causality principle and the dissipation-fluctuation theorem is
$$๐(๐ซ,t)=\epsilon _0_0^{\mathrm{}}d\tau \chi (๐ซ,\tau )๐(๐ซ,t\tau )+๐_\mathrm{N}(๐ซ,t),$$
(3.7)
where $`\chi (๐ซ,\tau )`$ is the dielectric susceptibility as a function of space and time, and $`๐_\mathrm{N}`$ is the (noise) polarization associated with absorption.
Substitution of this expression into Eq. (3.5) together with Fourier transformation converts this equation to
$$\underset{ยฏ}{๐}(๐ซ,\omega )=\epsilon _0\epsilon (๐ซ,\omega )\underset{ยฏ}{๐}(๐ซ,\omega )+\underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega ),$$
(3.8)
and thus
$$\underset{ยฏ}{๐}(๐ซ,\omega )=\epsilon _0\left[\epsilon (๐ซ,\omega )1\right]\underset{ยฏ}{๐}(๐ซ,\omega )+\underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega ),$$
(3.9)
where
$$\epsilon (๐ซ,\omega )=1+_0^{\mathrm{}}d\tau \chi (๐ซ,\tau )\mathrm{e}^{i\omega \tau }$$
(3.10)
is the (relative) permittivity, and the Maxwell equations (3.1) โ (3.4) read in the Fourier domain as<sup>4</sup><sup>4</sup>4Here and in the following the Fourier transform $`\underset{ยฏ}{F}(\omega )`$ of a real function $`F(t)`$ is defined according to the relation $`F(t)`$ $`=`$ $`F^{(+)}(t)`$ $`+`$ $`F^{()}(t)`$, where $`F^{(+)}(t)`$ $`=`$ $`_0^{\mathrm{}}d\omega \underset{ยฏ}{F}(\omega )\mathrm{e}^{i\omega t}`$ and $`F^{()}(t)`$ $`=`$ $`[F^{(+)}(t)]^{}`$.
$$\mathbf{}\underset{ยฏ}{๐}(๐ซ,\omega )=0,$$
(3.11)
$$\mathbf{}\times \underset{ยฏ}{๐}(๐ซ,\omega )i\omega \underset{ยฏ}{๐}(๐ซ,\omega )=0,$$
(3.12)
$$\epsilon _0\mathbf{}\epsilon (๐ซ,\omega )\underset{ยฏ}{๐}(๐ซ,\omega )=\underset{ยฏ}{\rho }_\mathrm{N}(๐ซ,\omega ),$$
(3.13)
$$\mathbf{}\times \underset{ยฏ}{๐}(๐ซ,\omega )+i\frac{\omega }{c^2}\epsilon (๐ซ,\omega )\underset{ยฏ}{๐}(๐ซ,\omega )=\mu _0\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ,\omega ).$$
(3.14)
Here we have introduced the (noise) charge density
$$\underset{ยฏ}{\rho }_\mathrm{N}(๐ซ,\omega )=\mathbf{}\underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega )$$
(3.15)
and the (noise) current density
$$\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ,\omega )=i\omega \underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega ),$$
(3.16)
which obey the continuity equation
$$\mathbf{}\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ,\omega )i\omega \underset{ยฏ}{\rho }_\mathrm{N}(๐ซ,\omega )=0.$$
(3.17)
According to Eq. (3.10), the permittivity $`\epsilon (๐ซ,\omega )`$ is a complex function of frequency,
$$\epsilon (๐ซ,\omega )=\epsilon _\mathrm{R}(๐ซ,\omega )+i\epsilon _\mathrm{I}(๐ซ,\omega ).$$
(3.18)
The real and imaginary parts, which are responsible for dispersion and absorption, respectively, are uniquely related to each other through the KramersโKronig relations
$$\epsilon _\mathrm{R}(๐ซ,\omega )1=\frac{๐ซ}{\pi }d\omega ^{}\frac{\epsilon _\mathrm{I}(๐ซ,\omega ^{})}{\omega ^{}\omega },$$
(3.19)
$$\epsilon _\mathrm{I}(๐ซ,\omega )=\frac{๐ซ}{\pi }d\omega ^{}\frac{\epsilon _\mathrm{R}(๐ซ,\omega ^{})1}{\omega ^{}\omega }$$
(3.20)
($`๐ซ`$, principal value). Further, $`\epsilon (๐ซ,\omega )`$ as a function of complex $`\omega `$ satisfies the relation
$$\epsilon (๐ซ,\omega ^{})=\epsilon ^{}(๐ซ,\omega )$$
(3.21)
and is holomorphic in the upper complex half-plane without zeros. In particular, it approaches unity in the high-frequency limit, i.e., $`\epsilon (๐ซ,\omega )`$ $``$ $`1`$ if $`|\omega |`$ $``$ $`\mathrm{}`$ .
The Maxwell equations (3.12) and (3.14) imply that $`\underset{ยฏ}{๐}(๐ซ,\omega )`$ obeys the partial differential equation
$$\mathbf{}\times \mathbf{}\times \underset{ยฏ}{๐}(๐ซ,\omega )\frac{\omega ^2}{c^2}\epsilon (๐ซ,\omega )\underset{ยฏ}{๐}(๐ซ,\omega )=i\omega \mu _0\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ,\omega ),$$
(3.22)
the solution of which can be represented in the form
$$\underset{ยฏ}{๐}(๐ซ,\omega )=i\omega \mu _0\mathrm{d}^3๐ซ^{}๐ฎ(๐ซ,๐ซ^{},\omega )\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ^{},\omega ),$$
(3.23)
where the Green tensor $`๐ฎ(๐ซ,๐ซ^{},\omega )`$ has to be determined from the equation
$$\mathbf{}\times \mathbf{}\times ๐ฎ(๐ซ,๐ซ^{},\omega )\frac{\omega ^2}{c^2}\epsilon (๐ซ,\omega )๐ฎ(๐ซ,๐ซ^{},\omega )=๐น(๐ซ๐ซ^{})$$
(3.24)
together with the boundary condition at infinity. In Cartesian coordinates, Eq.(3.24) reads
$$\left[\left(_i^r_k^r\delta _{ik}\mathrm{\Delta }^r\right)\delta _{ik}\frac{\omega ^2}{c^2}\epsilon (๐ซ,\omega )\right]G_{kj}(๐ซ,๐ซ^{},\omega )=\delta _{ij}\delta (๐ซ๐ซ^{})$$
(3.25)
($`_i^r`$ $`=`$ $`/x_i`$), where over repeated vector-component indices is summed. The Green tensor has the properties that
$$G_{ij}^{}(๐ซ,๐ซ^{},\omega )=G_{ij}(๐ซ,๐ซ^{},\omega ^{}),$$
(3.26)
$$G_{ji}(๐ซ^{},๐ซ,\omega )=G_{ij}(๐ซ,๐ซ^{},\omega ),$$
(3.27)
and
$$\mathrm{d}^3๐ฌ\frac{\omega ^2}{c^2}\epsilon _\mathrm{I}(๐ฌ,\omega )G_{ik}(๐ซ,๐ฌ,\omega )G_{jk}^{}(๐ซ^{},๐ฌ,\omega )=\mathrm{Im}G_{ij}(๐ซ,๐ซ^{},\omega ).$$
(3.28)
The property (3.26) is a direct consequence of the corresponding relation (3.21) for the permittivity. The reciprocity relation (3.27) and the integral relation (3.28) are proven in Appendix A.
The Fourier components of the magnetic induction, $`\underset{ยฏ}{๐}(๐ซ,\omega )`$, and the displacement field, $`\underset{ยฏ}{๐}(๐ซ,\omega )`$, are directly related to the Fourier components of the electric field, $`\underset{ยฏ}{๐}(๐ซ,\omega )`$,
$$\underset{ยฏ}{๐}(๐ซ,\omega )=(i\omega )^1\mathbf{}\times \underset{ยฏ}{๐}(๐ซ,\omega ),$$
(3.29)
$$\underset{ยฏ}{๐}(๐ซ,\omega )=(\mu _0\omega ^2)^1\mathbf{}\times \mathbf{}\times \underset{ยฏ}{๐}(๐ซ,\omega )$$
(3.30)
\[see Eqs. (3.12), (3.8), (3.16), and (3.22)\], and $`\underset{ยฏ}{๐}(๐ซ,\omega )`$ is determined, according to Eq. (3.23), by $`\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ,\omega )`$. The continuous set of (complex) fields $`\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ,\omega )`$ \[or equivalently, $`\underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega )`$\] can therefore be regarded as playing the role of the set of dynamical variables of the overall system composed of the electromagnetic field and the medium (including the dissipative system). For the following it is convenient to split off some factor from $`\underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega )`$ and to define the fundamental dynamical variables $`๐(๐ซ,\omega )`$ according to
$$\underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega )=i\sqrt{\frac{\mathrm{}\epsilon _0}{\pi }\epsilon _\mathrm{I}(๐ซ,\omega )}๐(๐ซ,\omega ).$$
(3.31)
### 3.2 Field quantization
The transition from classical to quantum theory now consists in the replacement of the classical fields $`๐(๐ซ,\omega )`$ and $`๐^{}(๐ซ,\omega )`$ by the operator-valued bosonic fields $`\widehat{๐}(๐ซ,\omega )`$ and $`\widehat{๐}^{}(๐ซ,\omega )`$, respectively, which are associated with the elementary excitations of the composed system within the framework of linear lightโmatter interaction. Thus the commutation relations are
$$[\widehat{f}_k(๐ซ,\omega ),\widehat{f}_k^{}^{}(๐ซ^{},\omega ^{})]=\delta _{kk^{}}\delta (๐ซ๐ซ^{})\delta (\omega \omega ^{}),$$
(3.32)
$$[\widehat{f}_k(๐ซ,\omega ),\widehat{f}_k^{}(๐ซ^{},\omega ^{})]=0,$$
(3.33)
and the Hamiltonian of the composed system is
$$\widehat{H}=\mathrm{d}^3๐ซ_0^{\mathrm{}}d\omega \mathrm{}\omega \widehat{๐}^{}(๐ซ,\omega )\widehat{๐}(๐ซ,\omega ).$$
(3.34)
Replacing $`\underset{ยฏ}{๐}(๐ซ,\omega )`$ \[Eq. (3.23)\], $`\underset{ยฏ}{๐}(๐ซ,\omega )`$ \[Eq. (3.29)\], and $`\underset{ยฏ}{๐}(๐ซ,\omega )`$ \[Eqs. (3.8), (3.30)\] by the quantum-mechanical operators, we find, on recalling Eqs. (3.16) and (3.31), that
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=i\sqrt{\frac{\mathrm{}}{\pi \epsilon _0}}\frac{\omega ^2}{c^2}\mathrm{d}^3๐ซ^{}\sqrt{\epsilon _\mathrm{I}(๐ซ^{},\omega )}๐ฎ(๐ซ,๐ซ^{},\omega )\widehat{๐}(๐ซ^{},\omega ),$$
(3.35)
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=(i\omega )^1\mathbf{}\times \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega ),$$
(3.36)
and
$`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=\epsilon _0\epsilon (๐ซ,\omega )\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\underset{ยฏ}{\overset{^}{๐}}_\mathrm{N}(๐ซ,\omega )`$ (3.37)
$`=(\mu _0\omega ^2)^1\mathbf{}\times \mathbf{}\times \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega ),`$
from which the electromagnetic field operators in the Schrรถdinger picture are obtained by integration over $`\omega `$:
$$\widehat{๐}(๐ซ)=_0^{\mathrm{}}d\omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.},$$
(3.38)
$$\widehat{๐}(๐ซ)=_0^{\mathrm{}}d\omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.},$$
(3.39)
and<sup>5</sup><sup>5</sup>5Here the longitudinal ($`๐
^{}`$) and transverse ($`๐
^{}`$) parts of a vector field $`๐
`$ are defined by $`๐
^{()}(๐ซ)`$ $`=`$ $`\mathrm{d}^3๐ซ^{}`$ $`๐น^{()}(๐ซ`$ $``$ $`๐ซ^{})๐
(๐ซ^{})`$, with $`๐น^{()}(๐ซ)`$ and $`๐น^{()}(๐ซ)`$ being respectively the longitudinal and transverse tensor-valued $`\delta `$-functions \[Eqs. (B.9) and (B.10)\]. Note that for bulk material the transverse polarization field $`\widehat{๐}^{}`$ $`=`$ $`\widehat{๐}`$ $``$ $`\epsilon _0\widehat{๐}^{}`$, with $`\widehat{๐}`$ and $`\widehat{๐}`$ being respectively given by Eqs. (3.38) and (3.40) together with Eq. (3.35) and (3.37) exactly corresponds to Eq. (2.24) together with Eq. (2.25).
$$\widehat{๐}(๐ซ)=\widehat{๐}^{}(๐ซ)=_0^{\mathrm{}}d\omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.}.$$
(3.40)
In this way, the electromagnetic field is expressed in terms of the classical Green tensor $`๐ฎ(๐ซ,๐ซ^{},\omega )`$ satisfying the generalized Helmholtz equation (3.24) and the continuum of the fundamental bosonic field variables $`\widehat{๐}(๐ซ,\omega )`$ \[and $`\widehat{๐}^{}(๐ซ,\omega )`$\]. All the information about the dielectric matter (such as its formation in space and its dispersive and absorptive properties) is contained \[via the permittivity $`\epsilon (๐ซ,\omega )`$\] in the Green tensor of the classical problem. Eqs. (3.38) โ (3.40), together with Eqs. (3.35) โ (3.37), can be considered as the generalization of the familiar mode decomposition.
A similar formalism which also starts from a causal relation between the polarization and the electric field strength is developed in . The auxiliary fields that are introduced there in order to construct a unitary time evolution in an enlarged Hilbert space can be shown to lead essentially to the field variables $`\widehat{๐}(๐ซ,\omega )`$ and $`\widehat{๐}^{}(๐ซ,\omega )`$ considered here. Thus, the representation of the electromagnetic field in corresponds to the Green function representation in Eqs. (3.35) โ (3.40).
The quantization scheme meets all the basic requirements of quantum electrodynamics. So it can be shown by using very general properties of the permittivity and the Green tensor that the electric and magnetic fields satisfy the correct (equal-time) commutation relations (see Appendix B)
$$[\widehat{E}_k(๐ซ),\widehat{E}_l(๐ซ^{})]=0=[\widehat{B}_k(๐ซ),\widehat{B}_l(๐ซ^{})],$$
(3.41)
$$[\epsilon _0\widehat{E}_k(๐ซ),\widehat{B}_l(๐ซ^{})]=i\mathrm{}ฯต_{klm}_m^r\delta (๐ซ๐ซ^{}).$$
(3.42)
Obviously, the electromagnetic field operators in the Heisenberg picture satisfy the Maxwell equations (3.1) โ (3.4), with the time derivative of any operator $`\widehat{Q}`$ being given by
$$\dot{\widehat{Q}}=(i\mathrm{})^1[\widehat{Q},\widehat{H}],$$
(3.43)
where $`\widehat{H}`$ is the Hamiltonian (3.34).
Let us briefly comment on the (zero-temperature) statistical implications of the quantization scheme. The vacuum expectation value of $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ is obviously zero whereas the fluctuation of $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ is not. From Eq. (3.35) together with the commutation relations (3.32) and (3.33) we derive, on using the integral relation (3.28),
$$0|\underset{ยฏ}{\overset{^}{E}}_k(๐ซ,\omega )\underset{ยฏ}{\overset{^}{E}}_l^{}(๐ซ^{},\omega ^{})|0=\frac{\mathrm{}\omega ^2}{\pi ฯต_0c^2}\mathrm{Im}G_{kl}(๐ซ,๐ซ^{},\omega )\delta (\omega \omega ^{}).$$
(3.44)
Equation (3.44) reveals that the fluctuation of the electromagnetic field is determined by the imaginary part of the Green tensor โ a result that is consistent with the dissipationโfluctuation theorem<sup>6</sup><sup>6</sup>6Note that the Green tensor plays the role of the response function of the electromagnetic field to an external perturbation. . Thus, the quantization scheme respects both the basic requirements of quantum theory (in terms of the correct commutation relations) and statistical physics (in terms of the dissipationโfluctuation theorem).
So far we have considered the electromagnetic field strengths. Instead, scalar ($`\widehat{\phi }`$) and vector ($`\widehat{๐}`$) potentials can be introduced and expressed in terms of the fundamental bosonic fields $`\widehat{๐}(๐ซ,\omega )`$ and $`\widehat{๐}^{}(๐ซ,\omega )`$. In particular, the potentials in the Coulomb gauge are defined by
$$\mathbf{}\widehat{\phi }(๐ซ)=\widehat{๐}^{}(๐ซ),$$
(3.45)
$$\widehat{๐}(๐ซ)=_0^{\mathrm{}}d\omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.},$$
(3.46)
where<sup>7</sup><sup>7</sup>7Note that for bulk material Eqs. (3.46) and (3.47) together with Eq. (3.35) exactly correspond to Eq. (2.23).
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=(i\omega )^1\underset{ยฏ}{\overset{^}{๐}}^{}(๐ซ,\omega ).$$
(3.47)
The canonically conjugated momentum field with respect to $`\widehat{๐}(๐ซ)`$ is
$$\widehat{๐ท}(๐ซ)=i\epsilon _0_0^{\mathrm{}}d\omega \omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.},$$
(3.48)
and it is not difficult to verify that $`\widehat{๐ท}`$ $`=`$ $`\epsilon _0\widehat{๐}^{}`$, $`\mathbf{}\times \widehat{๐}`$ $`=`$ $`\widehat{๐}`$, and $`\dot{\widehat{๐}}`$ $``$ $`\mathbf{}\widehat{\phi }`$ $`=`$ $`\widehat{๐}`$. In addition, $`\widehat{๐}`$ and $`\widehat{๐ท}`$ satisfy the well-known commutation relations (Appendix B)
$$[\widehat{A}_k(๐ซ),\widehat{A}_k^{}(๐ซ^{})]=0=[\widehat{\mathrm{\Pi }}_k(๐ซ),\widehat{\mathrm{\Pi }}_k^{}(๐ซ^{})],$$
(3.49)
$$[\widehat{A}_k(๐ซ),\widehat{\mathrm{\Pi }}_k^{}(๐ซ^{})]=i\mathrm{}\delta _{kk^{}}^{}(๐ซ๐ซ^{}).$$
(3.50)
#### 3.2.1 One-dimensional systems
Let us illustrate the main features of the concept for linearly polarized radiation propagating in the $`x`$ direction, which effectively reduces the system to one spatial dimension \[$`\widehat{๐}(๐ซ)`$ $``$ $`\widehat{A}_y(x)`$ $``$ $`\widehat{A}(x)`$, $`\widehat{๐ท}(๐ซ)`$ $``$ $`\widehat{\mathrm{\Pi }}_y(x)`$ $``$ $`\widehat{\mathrm{\Pi }}(x)`$, $`\widehat{๐}(๐ซ,\omega )`$ $``$ $`\widehat{f}(x,\omega )`$\]. According to Eqs. (3.46) and (3.47) together with Eq. (3.35), the operator of the vector potential is
$$\widehat{A}(x)=\sqrt{\frac{\mathrm{}}{\pi \epsilon _0๐}}_0^{\mathrm{}}d\omega dx^{}\frac{\omega }{c^2}\sqrt{\epsilon _\mathrm{I}(x^{},\omega )}G(x,x^{},\omega )\widehat{f}(x^{},\omega )+\text{H.c.},$$
(3.51)
and $`\widehat{\mathrm{\Pi }}`$ accordingly ($`๐`$, normalization area perpendicular to the $`x`$ direction). Here the Green function $`G(x,x^{},\omega )`$ satisfies the equation
$$\frac{^2}{x^2}G(x,x^{},\omega )\frac{\omega ^2}{c^2}\epsilon (x,\omega )G(x,x^{},\omega )=\delta (xx^{}).$$
(3.52)
In the simplest case when the spatial variation of the permittivity can be disregarded, then the solution of Eq. (3.52) that satisfies the boundary conditions at $`|x|,|x^{}|`$ $``$ $`\mathrm{}`$ is
$$G(x,x^{},\omega )=\left[2i\frac{\omega }{c}n(\omega )\right]^1\mathrm{exp}\left[i\frac{\omega }{c}n(\omega )|xx^{}|\right],$$
(3.53)
with $`n(\omega )`$ $`=`$ $`\sqrt{\epsilon (\omega )}`$ $`=`$ $`n_\mathrm{R}(\omega )`$ $`+`$ $`in_\mathrm{I}(\omega )`$ being the complex refractive index of the medium.
Substituting the Green function (3.53) into Eq. (3.51), we may rewrite the $`x^{}`$-integral to obtain
$`\widehat{A}(x)={\displaystyle _0^{\mathrm{}}}\mathrm{d}\omega \{\sqrt{{\displaystyle \frac{\mathrm{}}{4\pi \epsilon _0c\omega n_\mathrm{R}(\omega )๐}}}{\displaystyle \frac{n_\mathrm{R}(\omega )}{n(\omega )}}`$ (3.54)
$`\times [\mathrm{e}^{in_\mathrm{R}(\omega )\omega x/c}\widehat{a}_+(x,\omega )+\mathrm{e}^{in_\mathrm{R}(\omega )\omega x/c}\widehat{a}_{}(x,\omega )]+\text{H.c.}\},`$
where
$$\widehat{a}_\pm (x,\omega )=i\sqrt{2n_\mathrm{I}(\omega )\omega /c}\mathrm{e}^{n_\mathrm{I}(\omega )\omega x/c}_{\mathrm{}}^{\pm x}dx^{}\mathrm{e}^{in(\omega )\omega x^{}/c}\widehat{f}(\pm x^{},\omega ),$$
(3.55)
and from Eq. (3.32) it follows that
$$[\widehat{a}_\pm (x,\omega ),\widehat{a}_\pm ^{}(x^{},\omega ^{})]=\mathrm{e}^{n_\mathrm{I}(\omega )\omega |xx^{}|/c}\delta (\omega \omega ^{}).$$
(3.56)
Obviously, the space-dependent operators $`\widehat{a}_\pm (x,\omega )`$ describe the amplitudes of the damped monochromatic waves propagating to the right (subscript $`+`$) and left (subscript $``$), and from Eq. (3.55) it follows that $`\widehat{a}_\pm (x,\omega )`$ and $`\widehat{a}_\pm (x^{},\omega )`$ are related by (spatial) quantum Langevin equations ,
$$\frac{}{x}\widehat{a}_\pm (x,\omega )=n_\mathrm{I}(\omega )\frac{\omega }{c}\widehat{a}_\pm (x,\omega )+\widehat{F}_\pm (x,\omega ),$$
(3.57)
with
$$\widehat{F}_\pm (x,\omega )=\pm i\sqrt{2n_\mathrm{I}(\omega )\omega /c}\mathrm{e}^{in_\mathrm{R}(\omega )\omega x/c}\widehat{f}(x,\omega )$$
(3.58)
being the operator Langevin noise sources. In particular, when $`\widehat{f}(x^{\prime \prime },\omega )`$ $`=`$ $`0`$ for $`|x^{\prime \prime }`$ $``$ $`x|`$ $`<`$ $`|x`$ $``$ $`x^{}|`$ and $`|x^{\prime \prime }`$ $``$ $`x^{}|`$ $`<`$ $`|x`$ $``$ $`x^{}|`$, then
$$\widehat{a}_\pm (x,\omega )=\widehat{a}_\pm (x^{},\omega )\mathrm{exp}[n_\mathrm{I}(\omega )\omega |xx^{}|/c],\pm xx^{}0$$
(3.59)
for arbitrary $`\widehat{a}_\pm (x^{},\omega )`$.
Equation (3.54) is the extension of the familiar mode decomposition to absorbing media. Let us assume that in a frequency interval $`\mathrm{\Delta }\omega `$ the absorption is sufficiently small, so that for a chosen (finite) propagation interval $`|x`$ $``$ $`x^{}|`$ the condition $`n_\mathrm{I}(\omega )\omega |xx^{}|/c`$ $``$ $`1`$ holds. Then the amplitude operators $`\widehat{a}_\pm (x,\omega )`$ can be regarded as being independent of $`x`$ for that propagation distance and satisfying the ordinary Bose commutation relations, as is seen from Eqs. (3.55) and (3.56). In the chosen frequency and space intervals, Eq. (3.54) exactly reduces to the familiar expression obtained by mode expansion.
## 4 AtomโField Interaction
The interaction of the quantized electromagnetic field with atoms placed inside a dielectric-matter configuration or near dielectric bodies can strongly be influenced by the dielectric medium. A well-known example is the dependence of the rate of spontaneous decay of an excited atom on the properties of a dielectric environment (Sec. 6). In order to study such and related phenomena, the Hamiltonian (3.34) must be supplemented with the Hamiltonian of additional charged particles and their interaction energy with the medium-assisted electromagnetic field.
### 4.1 The minimal-coupling Hamiltonian
Applying the minimal-coupling scheme, we may write the total Hamiltonian in the form<sup>8</sup><sup>8</sup>8Here and in the following the subscripts A and M are introduced in order to distinguish between atom- and medium-assisted quantities.
$`\widehat{H}={\displaystyle \mathrm{d}^3๐ซ_0^{\mathrm{}}d\omega \mathrm{}\omega \widehat{๐}^{}(๐ซ,\omega )\widehat{๐}(๐ซ,\omega )}+{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right]^2`$ (4.1)
$`+\frac{1}{2}{\displaystyle \mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{A}(๐ซ)}+{\displaystyle \mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{M}(๐ซ)},`$
where $`\widehat{๐ซ}_\alpha `$ is the position operator and $`\widehat{๐ฉ}_\alpha `$ is the canonical momentum operator of the $`\alpha `$th (nonrelativistic) particle of charge $`q_\alpha `$ and mass $`m_\alpha `$. The Hamiltonian (4.1) consists of four terms. The first term is the energy of the electromagnetic field and the medium (including the dissipative system), as introduced in Eq. (3.34). The second term is the kinetic energy of the charged particles, and the third term is their Coulomb energy, where the corresponding scalar potential $`\widehat{\phi }_\mathrm{A}`$ is given by
$$\widehat{\phi }_\mathrm{A}(๐ซ)=\mathrm{d}^3๐ซ^{}\frac{\widehat{\rho }_\mathrm{A}(๐ซ^{})}{4\pi \epsilon _0|๐ซ๐ซ^{}|},$$
(4.2)
with
$$\widehat{\rho }_\mathrm{A}(๐ซ)=\underset{\alpha }{}q_\alpha \delta (๐ซ\widehat{๐ซ}_\alpha )$$
(4.3)
being the charge density of the particles. The last term is the Coulomb energy of interaction of the particles with the medium. From Eq. (4.1) it follows that the interaction Hamiltonian reads
$$\widehat{H}_{\mathrm{int}}=\underset{\alpha }{}\frac{q_\alpha }{m_\alpha }\left[\widehat{๐ฉ}_\alpha \frac{1}{2}q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right]\widehat{๐}(\widehat{๐ซ}_\alpha )+\mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{M}(๐ซ).$$
(4.4)
Note that in Eqs. (4.1) and (4.4), the vector potential $`\widehat{๐}`$ and the scalar potential $`\widehat{\phi }_\mathrm{M}`$, respectively, must be thought of as being expressed, on using Eq. (3.46) \[together with Eqs. (3.47) and (3.35)\] and Eq. (3.45) \[together with Eqs. (3.38) and (3.35)\], in terms of the fundamental fields $`\widehat{๐}(๐ซ,\omega )`$ \[and $`\widehat{๐}^{}(๐ซ,\omega )`$\].
In a straightforward but somewhat lengthy calculation (for an example, see Appendix C) it can be shown (by means of the commutation relations derived in Appendix B) that both the operator-valued Maxwell equations
$$\mathbf{}\widehat{๐}(๐ซ)=0,$$
(4.5)
$$\mathbf{}\times \widehat{๐}(๐ซ)+\dot{\widehat{๐}}(๐ซ)=0,$$
(4.6)
$$\mathbf{}\widehat{๐}(๐ซ)=\widehat{\rho }_\mathrm{A}(๐ซ),$$
(4.7)
$$\mathbf{}\times \widehat{๐}(๐ซ)\dot{\widehat{๐}}(๐ซ)=\widehat{๐ฃ}_\mathrm{A}(๐ซ),$$
(4.8)
where
$$\widehat{๐ฃ}_\mathrm{A}(๐ซ)=\frac{1}{2}\underset{\alpha }{}q_\alpha \left[\dot{\widehat{๐ซ}}_\alpha \delta (๐ซ\widehat{๐ซ}_\alpha )+\delta (๐ซ\widehat{๐ซ}_\alpha )\dot{\widehat{๐ซ}}_\alpha \right],$$
(4.9)
and the operator-valued Newtonian equations of motion
$$\dot{\widehat{๐ซ}}_\alpha =\frac{1}{m_\alpha }\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right],$$
(4.10)
$$m_\alpha \ddot{\widehat{๐ซ}}_\alpha =q_\alpha \left\{\widehat{๐}(\widehat{๐ซ}_\alpha )+\frac{1}{2}\left[\dot{\widehat{๐ซ}}_\alpha \times \widehat{๐}(\widehat{๐ซ}_\alpha )\widehat{๐}(\widehat{๐ซ}_\alpha )\times \dot{\widehat{๐ซ}}_\alpha \right]\right\}$$
(4.11)
are fulfilled. In Eqs. (4.6) โ (4.8), the (longitudinal part of the) electric field and the displacement field now contain \[compared with Eqs. (3.38) and (3.40)\] additional longitudinal components that result from the charge distribution $`\widehat{\rho }_\mathrm{A}(๐ซ)`$, i.e.,
$$\widehat{๐}(๐ซ)=\widehat{๐}_\mathrm{M}(๐ซ)\mathbf{}\widehat{\phi }_\mathrm{A}(๐ซ)=\left[_0^{\mathrm{}}d\omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.}\right]\mathbf{}\widehat{\phi }_\mathrm{A}(๐ซ),$$
(4.12)
$$\widehat{๐}(๐ซ)=\widehat{๐}_\mathrm{M}(๐ซ)\epsilon _0\mathbf{}\widehat{\phi }_\mathrm{A}(๐ซ)=\left[_0^{\mathrm{}}d\omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.}\right]\epsilon _0\mathbf{}\widehat{\phi }_\mathrm{A}(๐ซ),$$
(4.13)
with $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ and $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ being defined by Eqs. (3.35) and (3.37). The Maxwell equations (4.5) and (4.7), respectively, simply follow from the definition of $`\widehat{๐}(๐ซ)`$ \[Eq. (3.39) together with Eq. (3.36)\] and $`\widehat{๐}(๐ซ)`$ \[Eq. (4.13) together with Eqs. (3.37) and (4.2)\]. The Maxwell equations (4.6) and (4.8) are respectively the Heisenberg equations of motion of $`\widehat{๐}(๐ซ)`$ and $`\widehat{๐}(๐ซ)`$ according to Eq. (3.43), with the Hamiltonian being given by Eq. (4.1), and the Newtonian equations of motion (4.10) and (4.11) are respectively obtained from the Heisenberg equations of motion of $`\widehat{๐ซ}_\alpha `$ and $`\widehat{๐ฉ}_\alpha `$.
### 4.2 The multipolar-coupling Hamiltonian
In the interaction Hamiltonian (4.4) used in the minimal-coupling scheme the electromagnetic field is expressed in terms of the potentials. With regard to the interaction of the electromagnetic field with (localized) atomic systems (atoms, molecules etc.) the interaction energy is commonly desired to be treated in terms of the field strengths and the atomic polarization and magnetization. This can be achieved by means of a unitary transformation.
Let us consider an atomic system localized at position $`๐ซ_\mathrm{A}`$ and introduce the atomic polarization
$$\widehat{๐}_\mathrm{A}(๐ซ)=\underset{\alpha }{}q_\alpha \left(\widehat{๐ซ}_\alpha ๐ซ_\mathrm{A}\right)_0^1d\lambda \delta \left[๐ซ๐ซ_\mathrm{A}\lambda \left(\widehat{๐ซ}_\alpha ๐ซ_\mathrm{A}\right)\right],$$
(4.14)
so that the charge density (4.3) can be rewritten as
$$\widehat{\rho }_\mathrm{A}(๐ซ)=q_\mathrm{A}\delta (๐ซ\widehat{๐ซ}_\mathrm{A})\mathbf{}\widehat{๐}_\mathrm{A}(๐ซ),$$
(4.15)
with
$$q_\mathrm{A}=\underset{\alpha }{}q_\alpha $$
(4.16)
being the total charge of the atomic system. In order to perform the transition from the minimal-coupling scheme to the multipolar-coupling scheme, we apply to the variables the unitary operator
$$\widehat{U}=\mathrm{exp}\left[\frac{i}{\mathrm{}}\mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}(๐ซ)\right]$$
(4.17)
which is known as the PowerโZienau transformation . It is not difficult to prove that the following transformation rules are valid:
$$\widehat{๐ซ}_\alpha ^{}=\widehat{U}\widehat{๐ซ}_\alpha \widehat{U}^{}=\widehat{๐ซ}_\alpha ,$$
(4.18)
$`\widehat{๐ฉ}_\alpha ^{}=\widehat{U}\widehat{๐ฉ}_\alpha \widehat{U}^{}`$ (4.19)
$`=\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )q_\alpha {\displaystyle _0^1}d\lambda \lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\times \widehat{๐}\left[๐ซ_A+\lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\right],`$
$`\widehat{๐}^{}(๐ซ,\omega )=\widehat{U}\widehat{๐}(๐ซ,\omega )\widehat{U}^{}`$ (4.20)
$`=\widehat{๐}(๐ซ,\omega ){\displaystyle \frac{i}{\mathrm{}}}\sqrt{{\displaystyle \frac{\mathrm{}}{\pi \epsilon _0}}\epsilon _\mathrm{I}(๐ซ,\omega )}{\displaystyle \frac{\omega }{c^2}}{\displaystyle \mathrm{d}^3๐ซ^{}\widehat{๐}_\mathrm{A}^{}(๐ซ^{})๐ฎ^{}(๐ซ^{},๐ซ,\omega )}.`$
Employing equations (4.18) โ (4.20) and using Eqs. (3.28), (3.35), and (B.13), we can express the Hamiltonian $`\widehat{H}`$ in Eq. (4.1) in terms of the new variables $`\widehat{๐ซ}_\alpha ^{}`$ $`=`$ $`\widehat{๐ซ}_\alpha `$, $`\widehat{๐ฉ}_\alpha ^{}`$, and $`\widehat{๐}^{}(๐ซ,\omega )`$ in order to obtain the multipolar Hamiltonian
$`\widehat{H}={\displaystyle }\mathrm{d}^3๐ซ{\displaystyle _0^{\mathrm{}}}\mathrm{d}\omega \mathrm{}\omega \widehat{๐}^{}{}_{}{}^{}(๐ซ,\omega )\widehat{๐}^{}(๐ซ,\omega )`$ (4.21)
$`+{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}\left\{\widehat{๐ฉ}_\alpha ^{}+q_\alpha {\displaystyle _0^1}d\lambda \lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\times \widehat{๐}^{}\left[๐ซ_A+\lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\right]\right\}^2`$
$`+{\displaystyle \frac{1}{2}}{\displaystyle \mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{A}(๐ซ)}+{\displaystyle \frac{1}{2\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}^{}(๐ซ)\widehat{๐}_\mathrm{A}^{}(๐ซ)}`$
$`+{\displaystyle \mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{M}^{}(๐ซ)}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}^{}(๐ซ)\widehat{๐}_\mathrm{M}^{}(๐ซ)},`$
where the relations $`\widehat{๐}^{}`$ $`=`$ $`\widehat{๐}`$, $`\widehat{\phi }_\mathrm{M}^{}`$ $`=`$ $`\widehat{\phi }_\mathrm{M}`$ \[cf. Eq. (B.19)\], and
$$\widehat{๐}_\mathrm{M}^{}(๐ซ)=\widehat{๐}_\mathrm{M}^{}(๐ซ)+\frac{1}{\epsilon _0}\widehat{๐}_A^{}(๐ซ)$$
(4.22)
are valid.
In particular when the charged particles form a neutral atomic system ($`q_\mathrm{A}`$ $`=`$ $`0`$), then we may write, on integrating by parts and recalling that $`\widehat{๐}_{\mathrm{A}(\mathrm{M})}^{}`$ $`=`$ $`\widehat{๐}_{\mathrm{A}(\mathrm{M})}^{}/\epsilon _0`$,
$$\mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{A}(๐ซ)=\frac{1}{\epsilon _0}\mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}^{}(๐ซ)\widehat{๐}_\mathrm{A}^{}(๐ซ),$$
(4.23)
and
$`{\displaystyle \mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{M}(๐ซ)}={\displaystyle \frac{1}{\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}^{}(๐ซ)\widehat{๐}_\mathrm{M}^{}(๐ซ)}`$ (4.24)
$`={\displaystyle \frac{1}{\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}_\mathrm{M}(๐ซ)}{\displaystyle \frac{1}{\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}^{}(๐ซ)\widehat{๐}_\mathrm{M}^{}(๐ซ)}.`$
Combining Eqs. (4.21) โ (4.24) and taking into account that $`\widehat{๐}_\mathrm{M}^{}`$ $`=`$ $`\widehat{๐}_\mathrm{M}`$ \[cf. Eqs. (B.14) and (B.16)\], we may rewrite the multipolar Hamiltonian as
$`\widehat{H}={\displaystyle \mathrm{d}^3๐ซ_0^{\mathrm{}}d\omega \mathrm{}\omega \widehat{๐}^{}(๐ซ,\omega )\widehat{๐}^{}(๐ซ,\omega )}`$ (4.25)
$`+{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}\left\{\widehat{๐ฉ}_\alpha ^{}+q_\alpha {\displaystyle _0^1}d\lambda \lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\times \widehat{๐}^{}\left[๐ซ_A+\lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\right]\right\}^2`$
$`+{\displaystyle \frac{1}{2\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}_\mathrm{A}(๐ซ)}+{\displaystyle \frac{1}{\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}_\mathrm{M}^{}(๐ซ)}`$
$`{\displaystyle \frac{1}{\epsilon _0}}{\displaystyle }\mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}^{}{}_{}{}^{}(๐ซ),`$
where
$$\widehat{๐}^{}(๐ซ)=\widehat{๐}_\mathrm{M}^{}(๐ซ)=\epsilon _0\widehat{๐}_\mathrm{M}^{}(๐ซ)+\widehat{๐}_\mathrm{M}^{}(๐ซ).$$
(4.26)
From Eq. (4.25) the interaction Hamiltonian is seen to be
$`\widehat{H}_{\mathrm{int}}={\displaystyle \frac{1}{\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}_\mathrm{M}^{}(๐ซ)}{\displaystyle \frac{1}{\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}^{}(๐ซ)}`$ (4.27)
$`{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{q_\alpha }{2m_\alpha }}{\displaystyle _0^1}d\lambda \lambda \left\{\left[\left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\times \widehat{๐ฉ}_\alpha ^{}\right]\widehat{๐}^{}\left[๐ซ_A+\lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\right]+\text{H.c.}\right\}`$
$`+{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{q_\alpha ^2}{2m_\alpha }}\left\{{\displaystyle _0^1}d\lambda \lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\times \widehat{๐}^{}\left[๐ซ_A+\lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\right]\right\}^2.`$
The first term on the right-hand side in Eq. (4.27) is a contact term between the medium polarization and the polarization of the atomic system. The second term describes the interaction of the polarization of the atomic system with the transverse part of the overall displacement field \[cf. Eq. (4.26)\], and the last two terms refer to magnetic interactions.
So far we have transformed the dynamical variables but left unchanged the Hamiltonian. Instead the Hamiltonian can be transformed to obtain the new one
$$\widehat{}=\widehat{U}^{}\widehat{H}\widehat{U}.$$
(4.28)
Obviously, the new Hamiltonian expressed in terms of the new variables formally looks like the untransformed minimal-coupling Hamiltonian expressed in terms of the old variables. Hence expressing in the new Hamiltonian the new variables in terms of the old ones, we arrive at a multipolar-coupling Hamiltonian that (for a neutral atomic system) looks like the Hamiltonian given in Eq. (4.25), i.e.,
$`\widehat{}={\displaystyle \mathrm{d}^3๐ซ_0^{\mathrm{}}d\omega \mathrm{}\omega \widehat{๐}^{}(๐ซ,\omega )\widehat{๐}(๐ซ,\omega )}`$ (4.29)
$`+{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}\left\{\widehat{๐ฉ}_\alpha +q_\alpha {\displaystyle _0^1}d\lambda \lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\times \widehat{๐}\left[๐ซ_A+\lambda \left(\widehat{๐ซ}_\alpha ๐ซ_A\right)\right]\right\}^2`$
$`+{\displaystyle \frac{1}{2\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}_\mathrm{A}(๐ซ)}+{\displaystyle \frac{1}{\epsilon _0}}{\displaystyle \mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}_\mathrm{M}(๐ซ)}`$
$`{\displaystyle \frac{1}{\epsilon _0}}{\displaystyle }\mathrm{d}^3๐ซ\widehat{๐}_\mathrm{A}(๐ซ)\widehat{๐}{}_{}{}^{}(๐ซ).`$
In fact, the Hamiltonians in Eqs. (4.25) and (4.29) have different meanings. Since the expectation value of an observable associated with an operator $`\widehat{O}`$ must not change, the use of $`\widehat{}`$ necessarily requires a transformation of both the operator \[$`\widehat{O}`$ $``$ $`\widehat{U}^{}\widehat{O}\widehat{U}`$\] and the state \[$`\widehat{\varrho }`$ $``$ $`\widehat{U}^{}\widehat{\varrho }\widehat{U}`$, with $`\widehat{\varrho }`$ being the density operator\], so that
$$i\mathrm{}\frac{\mathrm{d}}{\mathrm{d}t}\widehat{O}=\mathrm{Tr}\left\{\widehat{\varrho }[\widehat{O},\widehat{H}]\right\}=\mathrm{Tr}\left\{\widehat{U}^{}\widehat{\varrho }\widehat{U}[\widehat{U}^{}\widehat{O}\widehat{U},\widehat{}]\right\}.$$
(4.30)
Moreover, in Eq. (4.25) the atomic polarization field couples to the transverse component of the overall displacement field, which also contains the transverse component of the atomic polarization, rather then the ordinary transverse displacement field in Eq. (4.29).
## 5 Inputโoutput coupling
The quantization scheme developed in Sec. 3 is best suited to study the inputโoutput behaviour of optical fields at dielectric devices, because both dispersion and absorption are exactly included in the resulting inputโoutput relations, and characteristic quantities such as the transmission and reflection coefficients of the setup are expressed in terms of its complex refractive-index profile . Inputโoutput relations are a very efficient description of the action of macroscopic bodies on radiation. In particular, they can advantageously be used to obtain the quantum statistics of the outgoing light from that of the incoming light, either in terms of radiation-field correlation functions or directly in terms of the density matrix. In what follows we will restrict our attention to four-port devices. The extension of the method to (higher-order) multiport-devices is straightforward.
### 5.1 Operator inputโoutput relations
Let us study the problem of propagation (in the $`x`$-direction) of quantized radiation through a dielectric plate that is the middle part of a three-layered planar structure as sketched in Fig. 1.
The setup can be characterized by a piecewise constant permittivity
$$\epsilon (x,\omega )=\underset{j=1}{\overset{3}{}}\lambda _j(x)\epsilon _j(\omega ),\lambda _j(x)=\{\begin{array}{cc}1\hfill & \text{if }x_{j1}<x<x_j\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array},$$
(5.1)
where $`\epsilon _j(\omega )`$ is the complex permittivity of the $`j`$th domain, and $`x_0`$ $``$ $`\mathrm{}`$, $`x_1`$ $`=`$ $`l/2`$, $`x_2`$ $`=`$ $`l/2`$, and $`x_3`$ $``$ $`+`$ $`\mathrm{}`$. The Green function $`G(x,x^{},\omega )`$ satisfies the partial differential equation (3.52) from which it follows that it can be decomposed into two parts<sup>9</sup><sup>9</sup>9For the calculation of Green tensors and a large variety of examples, see .
$$G(x,x^{},\omega )=\underset{j=1}{\overset{3}{}}\lambda _j(x)\lambda _j(x^{})G_j^{\mathrm{bulk}}(x,x^{},\omega )+R(x,x^{},\omega ),$$
(5.2)
where $`G_j^{\mathrm{bulk}}(x,x^{},\omega )`$ is the Green function of the bulk material as given by Eq. (3.53) (with $`\epsilon _j`$ in place of $`\epsilon `$), and the reflection term $`R(x,x^{},\omega )`$ is a solution to the homogeneous wave equation,
$$R(x,x^{},\omega )=\underset{j=1}{\overset{3}{}}\lambda _j(x)\left[C_{j+}(x^{},\omega )\mathrm{e}^{in_j(\omega )\omega x/c}+C_j(x^{},\omega )\mathrm{e}^{in_j(\omega )\omega x/c}\right],$$
(5.3)
where the coefficients $`C_{j\pm }(x^{},\omega )`$ are to be determined in such a way that continuity and differentiability at the surfaces of discontinuity at $`x`$ $`=`$ $`l/2`$ and $`x`$ $`=`$ $`l/2`$ are ensured. Combining Eqs. (3.51) and (5.2) \[together with Eq. (5.3)\], the vector potential $`\widehat{A}(x)`$ for the $`j`$th domain may be represented as, similar to Eq. (3.54),
$`\widehat{A}(x)={\displaystyle _0^{\mathrm{}}}\mathrm{d}\omega \{\sqrt{{\displaystyle \frac{\mathrm{}}{4\pi \epsilon _0c\omega n_{j\mathrm{R}}(\omega )๐}}}{\displaystyle \frac{n_{j\mathrm{R}}(\omega )}{n_j(\omega )}}`$ (5.4)
$`\times [\mathrm{e}^{in_{j\mathrm{R}}(\omega )\omega x/c}\widehat{a}_{j+}(x,\omega )+\mathrm{e}^{in_{j\mathrm{R}}(\omega )\omega x/c}\widehat{a}_j(x,\omega )]+\text{H.c.}\}`$
\[$`x_{j1}`$ $``$ $`x`$ $``$ $`x_j`$\], where the dependence on $`x`$ of the amplitude operators $`\widehat{a}_{j\pm }(x,\omega )`$ is governed by quantum Langevin equations of the type given in Eq. (3.57) together with Eq. (3.56).
The amplitude operators of the incoming fields are given according to Eq. (3.55) and satisfy the commutation relations<sup>10</sup><sup>10</sup>10Note that the amplitude operators of the field inside the plate (domain $`2`$) and the amplitude operators of the outgoing fields are not given according to Eq. (3.55) and thus do not satisfy commutation relations of this type in general.
$$[\widehat{a}_{1+}(x,\omega ),\widehat{a}_{1+}^{}(x^{},\omega ^{})]=\mathrm{e}^{n_{1\mathrm{I}}(\omega )\omega |xx^{}|/c}\delta (\omega \omega ^{}),$$
(5.5)
$$[\widehat{a}_3(x,\omega ),\widehat{a}_3^{}(x^{},\omega ^{})]=\mathrm{e}^{n_{3\mathrm{I}}(\omega )\omega |xx^{}|/c}\delta (\omega \omega ^{}),$$
(5.6)
$$[\widehat{a}_{1+}(x,\omega ),\widehat{a}_3^{}(x^{},\omega ^{})]=0$$
(5.7)
\[cf. Eq. (3.56)\]. Thus, the incoming fields from the left and right behave like the fields in the corresponding bulk dielectrics and may be regarded as independent variables. Further, it can be shown that the amplitude operators of the outgoing fields can be related to the amplitude operators of the incoming fields and appropriately chosen operators $`\widehat{g}_\pm (\omega )`$ of the plate as
$$\left(\genfrac{}{}{0pt}{}{\widehat{a}_1(\frac{1}{2}l,\omega )}{\widehat{a}_{3+}(\frac{1}{2}l,\omega )}\right)=๐(\omega )\left(\genfrac{}{}{0pt}{}{\widehat{a}_{1+}(\frac{1}{2}l,\omega )}{\widehat{a}_3(\frac{1}{2}l,\omega )}\right)+๐(\omega )\left(\genfrac{}{}{0pt}{}{\widehat{g}_+(\omega )}{\widehat{g}_{}(\omega )}\right),$$
(5.8)
where the 2$`\times `$2-matrices $`๐(\omega )`$ and $`๐(\omega )`$ are the characteristic transformation and absorption matrices of the plate expressed in terms of the thickness and the permittivity of the plate and the permittivities of the surrounding media, and the operators $`\widehat{g}_\pm (\omega )`$ read
$`\widehat{g}_\pm (\omega )=i\sqrt{{\displaystyle \frac{\omega }{2c\lambda _\pm (l,\omega )}}}\mathrm{e}^{in_2(\omega )\omega l/(2c)}`$ (5.9)
$`\times {\displaystyle _{l/2}^{l/2}}\mathrm{d}x^{}[\mathrm{e}^{in_2(\omega )\omega x^{}/c}\pm \mathrm{e}^{in_2(\omega )\omega x^{}/c}]\widehat{f}(x^{},\omega ),`$
with
$$\lambda _\pm (l,\omega )=\mathrm{e}^{n_{2\mathrm{I}}(\omega )\omega l/c}\left\{\frac{\mathrm{sinh}[n_{2\mathrm{I}}(\omega )\omega l/c]}{n_{2\mathrm{I}}(\omega )}\pm \frac{\mathrm{sin}[n_{2\mathrm{R}}(\omega )\omega l/c]}{n_{2\mathrm{R}}(\omega )}\right\}$$
(5.10)
(for details, see ). It is not difficult to prove, on recalling the basic commutation relations (3.32) and (3.33), that the commutation relations
$$[\widehat{g}_\pm (\omega ),\widehat{g}_\pm ^{}(\omega ^{})]=\delta (\omega \omega ^{}),$$
(5.11)
$$[\widehat{g}_\pm (\omega ),\widehat{g}_{}^{}(\omega ^{})]=0$$
(5.12)
are valid. Hence, the operators $`\widehat{g}_\pm (\omega )`$ and $`\widehat{g}_\pm ^{}(\omega )`$ are respectively annihilation and creation operators of bosonic excitations associated with the plate. Obviously, they commute with the amplitude operators of the incoming fields $`\widehat{a}_{1+}(x,\omega )`$ and $`\widehat{a}_3(x,\omega )`$ and are thus independent variables. It should be mentioned that the output amplitude operators $`\widehat{a}_1(x,\omega )`$, $`x`$ $``$ $`l/2`$, and $`\widehat{a}_{3+}(x,\omega )`$, $`x`$ $``$ $`l/2`$, can easily be obtained from $`\widehat{a}_1(l/2,\omega )`$ and $`\widehat{a}_{3+}(l/2,\omega )`$, respectively, by means of the corresponding solutions of the Langevin equations (3.57).
Whereas the input amplitude operators $`\widehat{a}_{1+}(l/2,\omega )`$, $`\widehat{a}_{1+}^{}(l/2,\omega )`$ and $`\widehat{a}_3(l/2,\omega )`$, $`\widehat{a}_3^{}(l/2,\omega )`$ satisfy ordinary bosonic commutation relations \[see Eqs. (5.5) โ (5.7)\], the amplitude operators of the outgoing fields, $`\widehat{a}_1(l/2,\omega )`$, $`\widehat{a}_1^{}(l/2,\omega )`$ and $`\widehat{a}_{3+}(l/2,\omega )`$, $`\widehat{a}_{3+}^{}(l/2,\omega )`$ do not, if the plate is surrounded by (absorbing) matter. When the plate is surrounded by vacuum, then the characteristic transformation and absorption matrices $`๐(\omega )`$ and $`๐(\omega )`$ fulfill the matrix relation<sup>11</sup><sup>11</sup>11Note that a unitary transformation together with rescaling can be applied to the output amplitude operators such that the new operators are bosonic and thus Eq. (5.13) can be assumed to be valid, without loss of generality.
$$๐(\omega )๐^+(\omega )+๐(\omega )๐^+(\omega )=๐$$
(5.13)
($`๐`$, unit matrix), and from Eq. (5.8) it then follows that the output amplitude operators also satisfy bosonic commutation relations. In fact, it can be shown that this is not only true at the very input and output ports of the plate but in the whole half-spaces on the left and the right. In this case both the input amplitude operators and the output amplitude operators can be regarded as being bosonic operators associated with ordinary monochromatic incoming and outgoing modes, respectively, and the matrices $`๐(\omega )`$ and $`๐(\omega )`$ read as \[$`n(\omega )n_2(\omega )`$\]
$$T_{11}(\omega )=T_{22}(\omega )=\mathrm{e}^{i\omega l/c}r(\omega )\left[1t_1(\omega )\mathrm{e}^{2in(\omega )\omega l/c}\vartheta (\omega )t_2(\omega )\right],$$
(5.14)
$$T_{12}(\omega )=T_{21}(\omega )=\mathrm{e}^{i\omega l/c}t_1(\omega )\mathrm{e}^{in(\omega )\omega l/c}\vartheta (\omega )t_2(\omega ),$$
(5.15)
$`A_{11}(\omega )=A_{21}(\omega )=\sqrt{n_\mathrm{I}(\omega )n_\mathrm{R}(\omega )}\mathrm{e}^{i\omega l/(2c)}t_1(\omega )\vartheta (\omega )`$ (5.16)
$`\times \sqrt{\lambda _+(l,\omega )}\left[1\mathrm{e}^{in(\omega )\omega l/c}r(\omega )\right],`$
$`A_{12}(\omega )=A_{22}(\omega )=\sqrt{n_\mathrm{I}(\omega )n_\mathrm{R}(\omega )}\mathrm{e}^{i\omega l/(2c)}t_1(\omega )\vartheta (\omega )`$ (5.17)
$`\times \sqrt{\lambda _{}(l,\omega )}\left[1+\mathrm{e}^{in(\omega )\omega l/c}r(\omega )\right].`$
Here,
$$r(\omega )=\frac{1n(\omega )}{1+n(\omega )}$$
(5.18)
and
$$t_1(\omega )=\frac{2}{1+n(\omega )},t_2(\omega )=\frac{2n(\omega )}{1+n(\omega )}$$
(5.19)
are the interface reflection and transmission coefficients, respectively, and the factor
$$\vartheta (\omega )=\left[1r^2(\omega )\mathrm{e}^{2in(\omega )\omega l/c}\right]^1$$
(5.20)
arises from multiple reflections inside the plate.
Operator inputโoutput relations of the type given in Eq. (5.8) are of course valid also for more complicated four-port devices such as multilayer plates. Obviously, the only difference consists in the actual expressions for the characteristic transformation and absorption matrices. For notational reasons it will be convenient to call the input amplitude operators $`\widehat{a}_i(\omega )`$ \[i.e., $`\widehat{a}_{1+}(l/2,\omega )`$ $``$ $`\widehat{a}_1(\omega )`$, $`\widehat{a}_3(l/2,\omega )`$ $``$ $`\widehat{a}_2(\omega )`$\], the output amplitude operators $`\widehat{b}_i(\omega )`$ \[i.e., $`\widehat{a}_1(l/2,\omega )`$ $``$ $`\widehat{b}_1(\omega )`$, $`\widehat{a}_{3+}(l/2,\omega )`$ $``$ $`\widehat{b}_2(\omega )`$\], the device operators $`\widehat{g}_i(\omega )`$ \[i.e., $`\widehat{g}_\pm (\omega )`$ $``$ $`\widehat{g}_i(\omega )`$\] (see Fig. 2), and to introduce the definitions
$$\widehat{๐}(\omega )=\left(\genfrac{}{}{0pt}{}{\widehat{a}_1(\omega )}{\widehat{a}_2(\omega )}\right),\widehat{๐}(\omega )=\left(\genfrac{}{}{0pt}{}{\widehat{b}_1(\omega )}{\widehat{b}_2(\omega )}\right),\widehat{๐ }(\omega )=\left(\genfrac{}{}{0pt}{}{\widehat{g}_1(\omega )}{\widehat{g}_2(\omega )}\right).$$
(5.21)
The operator input-output relations (5.8) at a four-port device can then be written in the compact form of
$$\widehat{๐}(\omega )=๐(\omega )\widehat{๐}(\omega )+๐(\omega )\widehat{๐ }(\omega ).$$
(5.22)
In what follows we assume that the characteristic transformation and absorption matrices $`๐(\omega )`$ and $`๐(\omega )`$, respectively, obey the equation (5.13).
### 5.2 Quantum-state transformation
The operator inputโoutput relations (5.22) can be used to calculate various moments and correlations of the outgoing fields in a straightforward way . The output operators are expressed in terms of the input operators and the device operators which for themselves act on the quantum state the incoming field and the device are prepared in. Instead the photonic operators may be left unchanged but the quantum state is transformed. This equivalent procedure is suitable in view of problems, such as the determination of the entanglement of the outgoing fields, where knowledge of the quantum state of the outgoing field as a whole is required. Hence, we are interested in a unitary transformation that transforms the input-state density operator $`\widehat{\varrho }_{\mathrm{in}}`$ (i.e., the density operator of the quantum state the incoming field and the device are prepared in) into an output-state density operator $`\widehat{\varrho }_{\mathrm{out}}`$ according to
$$\widehat{\varrho }_{\mathrm{out}}=\widehat{U}\widehat{\varrho }_{\mathrm{in}}\widehat{U}^{},$$
(5.23)
from which the density operator of the state the outgoing field is prepared in can be obtained by taking the trace with regard to the device variables.
Let us shortly digress to lossless devices and assume, for a moment, that $`๐(\omega )`$ $`=`$ $`0`$. It is clear from Eq. (5.13) that in this case the characteristic transformation matrix $`๐(\omega )`$ must be a unitary matrix and thus represents for each $`\omega `$ \[and $`\mathrm{det}๐(\omega )`$ $`=`$ $`1`$\] an element of the group SU(2) . For absorbing devices this can surely not be the case. Because of the coupling to the environment, we will definitely not be able to construct any unitary transformation that acts on the electromagnetic field operators alone. But we may look for one in the larger Hilbert space that comprises both the electromagnetic field and the device. Introducing some auxiliary field operators $`\widehat{๐ก}(\omega )`$ and defining the โfour-vectorsโ
$$\widehat{๐ถ}(\omega )=\left(\genfrac{}{}{0pt}{}{\widehat{๐}(\omega )}{\widehat{๐ }(\omega )}\right),\widehat{๐ท}(\omega )=\left(\genfrac{}{}{0pt}{}{\widehat{๐}(\omega )}{\widehat{๐ก}(\omega )}\right),$$
(5.24)
we may extend the inputโoutput relations (5.22) to the four-form of
$$\widehat{๐ท}(\omega )=๐ฒ(\omega )\widehat{๐ถ}(\omega ),$$
(5.25)
where $`๐ฒ(\omega )`$ is a unitary 4$`\times `$4-matrix, hence $`๐ฒ(\omega )๐ฒ^+(\omega )`$ $`=`$ $`๐ฐ`$. After separation of some phases from the matrices $`๐(\omega )`$ and $`๐(\omega )`$ and inclusion of them in the input operators, the matrix $`๐ฒ(\omega )`$ can be regarded (for each $`\omega `$) as an element of the group SU(4), and it can be expressed in terms of the matrices $`๐(\omega )`$ and $`๐(\omega )`$ as
$$๐ฒ(\omega )=(\begin{array}{cc}๐(\omega )& ๐(\omega )\\ ๐(\omega )๐^1(\omega )๐(\omega )& ๐(\omega )๐^1(\omega )๐(\omega )\end{array}),$$
(5.26)
where
$$๐(\omega )=\sqrt{๐(\omega )๐^+(\omega )}$$
(5.27)
and
$$๐(\omega )=\sqrt{๐(\omega )๐^+(\omega )}$$
(5.28)
are commuting positive Hermitian $`2\times 2`$-matrices. Note that $`๐^2(\omega )`$ $`+`$ $`๐^2(\omega )`$ $`=`$ $`๐`$.
The matrix transformation in Eq. (5.25) may be realized also as a unitary operator transformation
$$\widehat{๐ท}(\omega )=\widehat{U}^{}\widehat{๐ถ}(\omega )\widehat{U}=๐ฒ(\omega )\widehat{๐ถ}(\omega ),$$
(5.29)
where
$$\widehat{U}=\mathrm{exp}\left\{i_0^{\mathrm{}}d\omega \left[\widehat{๐ถ}^{}(\omega )\right]^T๐ฝ(\omega )\widehat{๐ถ}(\omega )\right\},$$
(5.30)
with the $`4\times 4`$-matrix $`๐ฝ(\omega )`$ being defined according to
$$\mathrm{exp}[i๐ฝ(\omega )]=๐ฒ(\omega ).$$
(5.31)
Obviously, $`\widehat{U}`$ is just the unitary operator that transforms $`\widehat{\varrho }_{\mathrm{in}}`$ into $`\widehat{\varrho }_{\mathrm{out}}`$ in Eq. (5.23). Since the input density operator can be regarded as being an operator functional of $`\widehat{๐ถ}(\omega )`$ and $`\widehat{๐ถ}^{}(\omega )`$, $`\widehat{\varrho }_{\mathrm{in}}`$ $`=`$ $`\widehat{\varrho }_{\mathrm{in}}[\widehat{๐ถ}(\omega ),\widehat{๐ถ}^{}(\omega )]`$, from Eqs. (5.23) and (5.29) it follows that the transformed density operator can be given by
$$\widehat{\varrho }_{\mathrm{out}}=\widehat{\varrho }_{\mathrm{in}}[๐ฒ^+(\omega )\widehat{๐ถ}(\omega ),๐ฒ^T(\omega )\widehat{๐ถ}^{}(\omega )].$$
(5.32)
Projecting $`\widehat{\varrho }_{\mathrm{out}}`$ onto the Hilbert space of the radiation field then yields the density operator of the outgoing fields
$$\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}=\mathrm{Tr}^{(\mathrm{D})}\left\{\widehat{\varrho }_{\mathrm{in}}[๐ฒ^+(\omega )\widehat{๐ถ}(\omega ),๐ฒ^T(\omega )\widehat{๐ถ}^{}(\omega )]\right\},$$
(5.33)
where $`\mathrm{Tr}^{(\mathrm{D})}`$ means the trace with respect to the device.<sup>12</sup><sup>12</sup>12The similarity to the usual open-systems approach to dissipation is not accidental. However, master (or related) equations, to which an open-systems theory would lead , are not required here, because the action of the environment (i.e., the device) is explicitly known. For chosen quantum state the device is initially prepared in, all the necessary information is contained in the characteristic transformation and absorption matrices $`๐(\omega )`$ and $`๐(\omega )`$, which in turn are determined by the Green function of the phenomenological Maxwell equations of the classical problem.
It is often useful and illustrative to describe quantum states in terms of phase-space functions such as the familiar $`s`$-parametrized phase-space functions (see, e.g., ). From Eq. (5.32) it follows that the $`s`$-parametrized phase-space functional $`P_{\mathrm{out}}[๐ถ(\omega );s]`$ that corresponds to $`\widehat{\varrho }_{\mathrm{out}}`$ is simply given by<sup>13</sup><sup>13</sup>13Since $`\omega `$ is continuous, $`P_{\mathrm{in}}[๐ถ(\omega );s]`$ and $`P_{\mathrm{out}}[๐ถ(\omega );s]`$ are functionals rather than functions.
$$P_{\mathrm{out}}[๐ถ(\omega );s]=P_{\mathrm{in}}[๐ฒ^+(\omega )๐ถ(\omega );s],$$
(5.34)
so that the phase-space functional of the outgoing radiation reads as
$$P_{\mathrm{out}}^{(\mathrm{F})}[๐ถ(\omega );s]=๐๐ P_{\mathrm{out}}[๐ถ(\omega );s]=๐๐ P_{\mathrm{in}}[๐ฒ^+(\omega )๐ถ(\omega );s],$$
(5.35)
where the functional integration (notation $`๐๐ `$) is taken over the continua of the complex phase-space variables $`g_1(\omega )`$ and $`g_2(\omega )`$ of the dielectric device. In Eq. (5.34) we have used the fact that application of the unitary transformation under consideration implies preservation of operator ordering, i.e., the annihilation and creation operators are not mixed by the quantum-state transformation. It should be pointed out that this is not the case for amplifying devices .
For practical purposes a description of the incoming and outgoing radiation in terms of discrete (quasi-monochromatic) modes is frequently preferred to be used. For this, we divide the frequency axis into sufficiently small intervals<sup>14</sup><sup>14</sup>14The matrices $`๐(\omega )`$ and $`๐(\omega )`$ must not change substantially over an interval. of mid-frequencies $`\omega _m`$ and widths $`\mathrm{\Delta }\omega _m`$ and define the discrete photonic input operators
$$\widehat{๐ถ}_m=\frac{1}{\sqrt{\mathrm{\Delta }\omega _m}}_{\mathrm{\Delta }\omega _m}d\omega \widehat{๐ถ}(\omega ),$$
(5.36)
and the discrete photonic output operators $`\widehat{๐ท}_m`$ accordingly. Then we assign to each pair of operators $`\widehat{๐ถ}_m`$ and $`\widehat{๐ท}_m`$ the inputโoutput relation (5.25) with the 4$`\times `$4-matrix $`๐ฒ_m=๐ฒ(\omega _m)`$, and, according to Eq. (5.30), the unitary operator $`\widehat{U}`$ then reads as
$$\widehat{U}=\underset{m}{}\widehat{U}_m,$$
(5.37)
where
$$\widehat{U}_m=\mathrm{exp}\left[i\left(\widehat{๐ถ}_m^{}\right)^T๐ฝ_m\widehat{๐ถ}_m\right]$$
(5.38)
with $`๐ฝ_m`$ $`=`$ $`๐ฝ(\omega _m)`$. In particular, the functional integral in Eq. (5.35) becomes an ordinary multiple integral.
#### 5.2.1 Examples
Let us restrict our attention to (quasi-)monochromatic fields, so that it is sufficient to consider only a single frequency component $`\omega _m`$. Suppose the incoming field and the device are prepared in coherent states<sup>15</sup><sup>15</sup>15For notational convenience we omit the subscript $`m`$.
$$|\psi _{\mathrm{in}}=|๐ธ=\mathrm{exp}\left(๐ธ^T\widehat{๐ถ}^{}๐ธ^+\widehat{๐ถ}\right)|0,๐ธ=\left(\genfrac{}{}{0pt}{}{๐}{๐}\right),$$
(5.39)
with $`c_j`$ and $`d_j`$ ($`j`$ $`=`$ $`1,2`$), respectively, being the coherent-state amplitudes of the input fields and the device. Application of Eq. (5.32) yields $`|\psi _{\mathrm{out}}`$ $`=`$ $`|๐ธ^{}`$, with $`๐ธ^{}`$ $`=`$ $`๐ฒ`$$`๐ธ`$ in place of $`๐ธ`$ in Eq. (5.39), and it follows, on applying Eq. (5.33), that the outgoing fields are prepared in coherent states, i.e.,
$$\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}=|๐^{}๐^{}|,๐^{}=\mathrm{๐๐}+\mathrm{๐๐}.$$
(5.40)
Thus, the coherent-state amplitudes $`c_1^{}`$ and $`c_2^{}`$ of the outgoing fields are not only determined by the characteristic transformation matrix $`๐`$ but also by the absorption matrix $`๐`$ via the coherent-state amplitudes of the device.
Next let us consider the case where the field in one of the two input channels is prepared in an $`n`$-photon Fock state and the field in the other input channel and the device are left in vacuum, i.e., $`|\psi _{\mathrm{in}}`$ $`=`$ $`|n000`$. The Wigner function of the input state reads<sup>16</sup><sup>16</sup>16$`W_{\mathrm{in}}(๐ถ)`$ $`=`$ $`W_n(a_1)W_0(a_2)W_0(g_1)W_0(g_2)`$, where $`W_k(a)`$ $`=`$ $`(2/\pi )(1)^kL_k(4|a|^2)\mathrm{e}^{2|a|^2}`$ \[$`L_k(z)`$, Laguerre polynomial\] is the Wigner function of a $`k`$-quanta Fock state (see, e.g., ).
$$W_{\mathrm{in}}(๐ถ)=(1)^n\left(\frac{2}{\pi }\right)^4L_n\left(4|a_1|^2\right)\mathrm{exp}\left(2|๐ถ|^2\right).$$
(5.41)
Applying Eqs. (5.34) and (5.35) and integrating over the phase space of one outgoing field, the Wigner function of the field in the $`j`$th output channel is derived to be
$$W_{\mathrm{out}}^{(\mathrm{F})}\left(a_j\right)=\underset{k=0}{\overset{n}{}}\left(\genfrac{}{}{0pt}{}{n}{k}\right)|T_{j1}|^{2k}\left(1|T_{j1}|^2\right)^{nk}W_k(a_j).$$
(5.42)
Obviously, Eq. (5.42) does not only hold for the Wigner function but for any $`s`$-parametrized phase-space function, and hence the corresponding density operator reads in the Fock basis as
$$\widehat{\varrho }_{\mathrm{out}j}^{(\mathrm{F})}=\underset{k=0}{\overset{n}{}}\left(\genfrac{}{}{0pt}{}{n}{k}\right)|T_{j1}|^{2k}\left(1|T_{j1}|^2\right)^{nk}|kk|.$$
(5.43)
Note that the quantum state the outgoing field is prepared in contains only Fock states up to the photon number $`n`$ of the input Fock state. That is a direct consequence of the action of the compact group SU(4) which leaves the total number of quanta unchanged. Furthermore, the Fock state with the highest photon number (i.e., the input state) appears in the output state with a weight $`|T_{j1}|^{2n}`$, that is, the probability of finding the same number of photons after transmission through (reflection at) an absorbing device decreases as $`|T_{j1}|^{2n}`$. In fact, for a device in the ground state, as it is the case here, there is some classical reasoning explaining this result. One could imagine that each of the $`n`$ photons โfeelsโ the effect of a transmission (reflection) coefficient smaller than unity, which for $`n`$ photons amounts to the $`n`$th power of the transmission (reflection) coefficient. This reasoning fails when the device is prepared in anything else then the ground state.
Finally, let us assume that the field in one of the two input channels is prepared in a Schrรถdinger-cat state and the field in the other input channel and the device are left in vacuum, i.e.,
$$|\psi _{\mathrm{in}}=\frac{1}{\sqrt{N}}\left(|\gamma +|\gamma \right)|000,$$
(5.44)
where $`|\gamma `$ is a coherent state, and $`N`$ $`=`$ $`2[1`$ $`+`$ $`\mathrm{exp}(2|\gamma |^2)]`$ the proper normalization factor. Using Eq. (5.40) with $`c_1`$ $`=`$ $`\pm \gamma `$, $`c_2`$ $`=`$ $`d_1`$ $`=`$ $`d_2`$ $`=`$ $`0`$, it is not difficult to derive the density operator of the field in the $`j`$th output channel. The result is
$`\widehat{\varrho }_{\mathrm{out}j}^{(\mathrm{F})}={\displaystyle \frac{1}{N}}\{|\gamma T_{j1}\gamma T_{j1}|+|\gamma T_{j1}\gamma T_{j1}|`$ (5.45)
$`+(|\gamma T_{j1}\gamma T_{j1}|+|\gamma T_{j1}\gamma T_{j1}|)\mathrm{exp}[2|\gamma |^2(1|T_{j1}|^2)]\}.`$
Whereas the two peaks decay as $`|T_{j1}|^2`$, the quantum interference, in contrast, decays exponentially as $`|T_{j1}|^2\mathrm{exp}[2|\gamma |^2(1|T_{j1}|^2)]`$. The larger the mean number of photons $`\widehat{n}`$ $`=`$ $`|\gamma |^2\mathrm{tanh}|\gamma |^2`$ becomes, the faster is the decay. Let us consider, e.g., the transmitted Schrรถdinger cat ($`j`$ $`=`$ $`2`$). From Eq. (5.13) it follows that the squares of the absolute values of the transmission coefficient $`|T_{21}|^2`$, the reflection coefficient $`|T_{22}|^2`$, and the absorption coefficients $`|A_{21}|^2`$ and $`|A_{22}|^2`$ are related to each other as $`1`$ $``$ $`|T_{21}|^2`$ $`=`$ $`|T_{22}|^2`$ $`+`$ $`|A_{21}|^2`$ $`+`$ $`|A_{22}|^2`$, so that $`\mathrm{exp}[2|\gamma |^2(1|T_{j1}|^2)]`$ $`=`$ $`\mathrm{exp}[2|\gamma |^2|T_{22}|^2)`$ $`\mathrm{exp}[2|\gamma |^2(|A_{21}|^2`$ $`+`$ $`|A_{22}|^2)]`$. The terms $`\mathrm{exp}(2|\gamma |^2|T_{22}|^2)`$ and $`\mathrm{exp}[2|\gamma |^2(|A_{21}|^2`$ $`+`$ $`|A_{22}|^2)]`$ then respectively describe the decoherence associated with the losses owing to reflection and absorption.
#### 5.2.2 Entanglement degradation
Quantum information processing such as quantum teleportation and quantum cryptography is essentially based on entanglement, which can be regarded as being the nonclassical contribution to the overall correlation between two parts of a system. In particular, when two fields that are prepared in nonclassical states are superimposed by a four-port device, then the two outgoing fields are prepared in an entangled state in general. Entanglement of a bipartite quantum state $`\widehat{\varrho }`$ is commonly quantified by measures which fulfill some basic requirements as non-negativity (being zero only for separable states), invariance under local unitary transformations of the subsystems, and non-increase under arbitrary positive trace-preserving maps . Further, the reduced von Neumann entropy should be recovered for pure states which in addition gives a normalization condition for the measure. So far, the distance $`E(\widehat{\varrho })`$ of $`\widehat{\varrho }`$ to the set $`๐ฎ`$ of all separable quantum states $`\widehat{\sigma }`$<sup>17</sup><sup>17</sup>17A quantum state $`\widehat{\sigma }`$ of a system that consists of two subsystems $`A`$ and $`B`$ is called separable if $`\widehat{\sigma }`$ $`=`$ $`_ip_i\widehat{\sigma }_i^{(A)}\widehat{\sigma }_i^{(B)}`$, where $`\widehat{\sigma }_i^{(A)}`$ and $`\widehat{\sigma }_i^{(B)}`$ represent quantum states of the subsystems, and $`_ip_i`$ $`=`$ $`1`$. which is measured by means of the relative entropy has been proven to be one measure satisfying all the given conditions . Thus,
$$E(\widehat{\varrho })=\underset{\widehat{\sigma }๐ฎ}{\mathrm{min}}\mathrm{Tr}\left[\widehat{\varrho }\left(\mathrm{ln}\widehat{\varrho }\mathrm{ln}\widehat{\sigma }\right)\right].$$
(5.46)
Let us consider the entanglement degradation that is observed when two fields that are prepared in a Bell basis state<sup>18</sup><sup>18</sup>18For the entanglement degradation of a two-mode squeezed vacuum, see .
$$|\mathrm{\Psi }^\pm =\frac{1}{\sqrt{2}}\left(|01\pm |10\right)$$
(5.47)
or
$$|\mathrm{\Phi }^\pm =\frac{1}{\sqrt{2}}\left(|00\pm |11\right)$$
(5.48)
are transmitted through absorbing four-port devices prepared in the ground state.<sup>19</sup><sup>19</sup>19For an application to photon tunneling through absorbing dielectric barriers, see . Note that the entanglement of each of the Bell basis states (5.47) and (5.48) is equal to $`\mathrm{ln}2`$ (sometimes also named $`1`$ bit). Applying Eqs. (5.32) and (5.33) to the input states $`|\psi _{\mathrm{in}}`$ $`=`$ $`|\mathrm{\Psi }^\pm |0_\mathrm{D}`$ and $`|\psi _{\mathrm{in}}`$ $`=`$ $`|\mathrm{\Phi }^\pm |0_\mathrm{D}`$ ($`|0_\mathrm{D}`$, ground state of the two devices), we easily find that the outgoing fields are prepared in the quantum states
$`\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}(|\mathrm{\Psi }^\pm )=\frac{1}{2}\left[\left(2|T_1|^2|T_2|^2\right)|0000|\right]`$ (5.49)
$`+\frac{1}{2}\left(T_2|01\pm T_1|10\right)\left(T_2^{}01|\pm T_1^{}10|\right),`$
$`\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}(|\mathrm{\Phi }^\pm )=\frac{1}{2}[(1|T_1|^2)(1|T_2|^2)|0000|`$ (5.50)
$`+|T_1|^2(1|T_2|^2)|1010|+|T_2|^2(1|T_1|^2)|0101|]`$
$`+\frac{1}{2}\left(|00\pm T_1T_2|11\right)\left(00|\pm (T_1T_2)^{}11|\right).`$
Here and in the following the notation $`(T_l)_{11}`$ $`=`$ $`(T_l)_{22}`$ $``$ $`R_l`$ and $`(T_l)_{12}`$ $`=`$ $`(T_l)_{21}`$ $``$ $`T_l`$ for the elements of the characteristic transformation matrix $`๐_l`$ of the $`l`$th four-port device is used ($`l`$ $`=`$ $`1,2`$) \[cf. Eqs. (5.14) and (5.15)\]. The entanglement contained in the quantum states of the outgoing fields, $`\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}(|\mathrm{\Psi }^\pm )`$ and $`\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}(|\mathrm{\Phi }^\pm )`$, can be estimated by using the convexity property of the relative entropy.<sup>20</sup><sup>20</sup>20$`E(_ip_i\widehat{\varrho }_i)`$ $``$ $`_ip_iE(\widehat{\varrho }_i)`$, where $`_ip_i`$ $`=`$ $`1`$ . The partition into separable states and a single pure state is not unique. Is has been proven, however, that for a pair of spin-$`\frac{1}{2}`$ parties there exists a unique (sometimes called optimal) decomposition such that the weight of the separable state is maximized . The reduced von Neumann entropy of the extracted pure state with corresponding minimal weight is then exactly the amount of entanglement, hence the inequality reduces to an equality. In Eqs. (5.49) and (5.50) the output state is written as a sum of separable states and a single pure state. Since separable states have zero entanglement by definition, the entanglement of the whole state is bounded from above by the reduced von Neumann entropy of the respective pure states. If we assume equal transmission coefficients of the two devices ($`T_1`$ $`=`$ $`T_2`$ $`=`$ $`T`$), the bounds are given according to
$$E\left[\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}(|\mathrm{\Psi }^\pm )\right]|T|^2\mathrm{ln}2$$
(5.51)
and
$$E\left[\widehat{\varrho }_{\mathrm{out}}^{(\mathrm{F})}(|\mathrm{\Phi }^\pm )\right]\frac{1}{2}\left[\left(1+|T|^4\right)\mathrm{ln}\left(1+|T|^4\right)|T|^4\mathrm{ln}|T|^4\right].$$
(5.52)
Suppose the two fields are transmitted through (equal) optical fibres with perfect input coupling ($`R`$ $`=`$ $`0`$), so that the transmission coefficient $`T`$ may be given by
$$T=\mathrm{e}^{in\omega l/c}=\mathrm{e}^{in_\mathrm{R}\omega l/c}\mathrm{e}^{l/L},$$
(5.53)
with $`l`$ and $`L`$ $`=`$ $`c/(n_\mathrm{I}\omega )`$ being respectively the propagation length through the fibres and the absorption length of the fibres. Substituting of this expression into the inequalities (5.51) and (5.52) yields the dependence on $`l`$ of the bounds of entanglement.
The exact dependence on $`l`$ of the entanglement degradation calculated by applying Eq. (5.46) is shown in Fig. 3. One observes that the states $`|\mathrm{\Phi }^\pm `$ decay faster than the states $`|\mathrm{\Psi }^\pm `$. Since the device has been left in the ground state, we can again use some classical reasoning to explain this behaviour. When the two fields are initially prepared in a state $`|\mathrm{\Psi }^\pm `$, then only a single photon (i.e., either the photon from the first field or the photon from the second one) is effectively subject to absorption. By contrast, the two photons get effected simultaneously, if the initial state is a state $`|\mathrm{\Phi }^\pm `$.
## 6 Spontaneous decay
Spontaneous emission is a prime example of the action of ground-state fluctuations on physically measurable processes. Einstein already pointed out that, in order to obtain the Planck radiation law, a process as spontaneous emission must necessarily be included in the theory of atomic decay. Later on, the radiation properties of an excited atom located in free space have been a subject of many studies. In particular, the rate of spontaneous emission of an excited (two-level) atom in free space is given by (the famous Einstein $`A`$-coefficient)
$$\mathrm{\Gamma }_0=\frac{\omega _\mathrm{A}^3d^2}{3\pi \mathrm{}\epsilon _0c^3},$$
(6.1)
where $`d`$ is the absolute value of the matrix element $`๐`$ $`=`$ $`l|\widehat{๐}_\mathrm{A}|u`$ of the atomic dipole operator $`\widehat{๐_\mathrm{A}}`$ \[Eq. (6.6)\] between the upper state $`|u`$ and the lower state $`|l`$, and $`\omega _\mathrm{A}`$ is the corresponding atomic transition frequency. When the atom is surrounded by (dielectric) matter, then the ground state felt by the radiating atom is changed and thus the rate formula (6.1) must be corrected<sup>21</sup><sup>21</sup>21In the strong-coupling regime the decay becomes non-exponential and cannot be described by a rate (Sec. 6.4). .
### 6.1 Time dependence of the atomโfield system
#### 6.1.1 Hamiltonian
Let us assume that the surrounding matter can be regarded as being a dielectric of given complex permittivity. Applying the minimal-coupling scheme (Sec. 4.1), we may decompose the Hamiltonian \[Eq. (4.1)\] of the coupled system consisting of an atom and the medium-assisted electromagnetic field as
$$\widehat{H}=\widehat{H}_\mathrm{A}+\widehat{H}_\mathrm{M}+\widehat{H}_{\mathrm{int}},$$
(6.2)
where
$$\widehat{H}_\mathrm{A}=\underset{\alpha }{}\frac{\widehat{๐ฉ}_\alpha ^2}{2m_\alpha }+\frac{1}{2}\mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{A}(๐ซ)$$
(6.3)
is the Hamiltonian of the atom,
$$\widehat{H}_\mathrm{M}=\mathrm{d}^3๐ซ_0^{\mathrm{}}d\omega \mathrm{}\omega \widehat{๐}^{}(๐ซ,\omega )\widehat{๐}(๐ซ,\omega )$$
(6.4)
is the Hamiltonian of the electromagnetic field and the dielectric matter, and
$$\widehat{H}_{\mathrm{int}}=\underset{\alpha }{}\frac{q_\alpha }{m_\alpha }\widehat{๐ฉ}_\alpha \widehat{๐}(๐ซ_\alpha )+\mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{M}(๐ซ)$$
(6.5)
is the interaction energy. Here we have omitted the $`\widehat{๐}^2`$ term. In the electric-dipole approximation, the first term on the right-hand side of Eq. (6.5) simplifies to
$$\underset{\alpha }{}\frac{q_\alpha }{m_\alpha }\widehat{๐ฉ}_\alpha \widehat{๐}(๐ซ_\alpha )=\frac{1}{i\mathrm{}}[\widehat{๐}_\mathrm{A},\widehat{H}_\mathrm{A}]\widehat{๐}(๐ซ_\mathrm{A}),$$
(6.6)
where
$$\widehat{๐}_\mathrm{A}=\underset{\alpha }{}q_\alpha \widehat{๐ซ}_\alpha $$
(6.7)
is the atomic dipole operator. Restricting our attention to a two-level system, the atomic Hamiltonian $`\widehat{H}_\mathrm{A}`$ reduces to
$$\widehat{H}_\mathrm{A}=\mathrm{}\omega _u|uu|+\mathrm{}\omega _l|ll|=\frac{1}{2}\mathrm{}\omega _\mathrm{A}\widehat{\sigma }_z+\text{const.},$$
(6.8)
where $`\widehat{\sigma }_z`$ $`=`$ $`|uu|`$ $``$ $`|ll|`$. Thus we may further simplify Eq. (6.6) to obtain, on applying the rotating wave approximation and recalling Eqs. (3.46) and (3.47),
$$\underset{\alpha }{}\frac{q_\alpha }{m_\alpha }\widehat{๐ฉ}_\alpha \widehat{๐}(๐ซ_\alpha )=\widehat{\sigma }^{}\widehat{๐}_\mathrm{M}^{(+)}(๐ซ_\mathrm{A})๐+\text{H.c.}$$
(6.9)
($`๐`$ real), with $`\widehat{\sigma }`$ being defined by $`\widehat{\sigma }`$ $`=`$ $`|lu|`$.
For a neutral atom, the atomic polarization defined by Eq. (4.14) reduces in the dipole approximation to
$$\widehat{๐}_\mathrm{A}(๐ซ)=\widehat{๐}_\mathrm{A}\delta (๐ซ๐ซ_\mathrm{A}),$$
(6.10)
and thus application of Eq. (4.24) yields ($`\widehat{๐}_\mathrm{M}^{}`$ $`=`$ $`\widehat{๐}_\mathrm{M}^{}/\epsilon _0`$)
$$\mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{M}(๐ซ)=\widehat{๐}_\mathrm{A}\widehat{๐}_\mathrm{M}^{}(๐ซ_\mathrm{A}),$$
(6.11)
which for a two-level atom in the rotating wave approximation reads
$$\mathrm{d}^3๐ซ\widehat{\rho }_\mathrm{A}(๐ซ)\widehat{\phi }_\mathrm{M}(๐ซ)=\widehat{\sigma }^{}\widehat{๐}_\mathrm{M}^{(+)}(๐ซ_\mathrm{A})๐+\text{H.c.}.$$
(6.12)
Eqs. (6.9) and (6.12) yield
$$\widehat{H}_{\mathrm{int}}=\widehat{\sigma }^{}\widehat{๐}_\mathrm{M}^{(+)}(๐ซ_\mathrm{A})๐+\text{H.c.}.$$
(6.13)
#### 6.1.2 Heisenberg picture
From the Hamiltonian given in Eq. (6.2) together with Eqs. (6.4), (6.8), and (6.13), the Heisenberg equations of motion read
$$\dot{\widehat{\sigma }}_z=\frac{2i}{\mathrm{}}\widehat{\sigma }^{}\widehat{๐}_\mathrm{M}^{(+)}(๐ซ_\mathrm{A})๐+\text{H.c.},$$
(6.14)
$$\dot{\widehat{\sigma }}{}_{}{}^{}=i\omega _\mathrm{A}\widehat{\sigma }^{}+\frac{i}{\mathrm{}}\widehat{๐}_\mathrm{M}^{()}(๐ซ_\mathrm{A})๐\widehat{\sigma }_z,$$
(6.15)
$$\dot{\widehat{๐}}(๐ซ,\omega )=i\omega \widehat{๐}(๐ซ,\omega )+\frac{\omega ^2}{c^2}\sqrt{\frac{\epsilon _\mathrm{I}(๐ซ,\omega )}{\mathrm{}\pi \epsilon _0}}๐ฎ^{}(๐ซ,๐ซ_\mathrm{A},\omega )๐\widehat{\sigma }.$$
(6.16)
Substituting in the expression for the electric field strength \[Eq. (3.38) together with Eq. (3.35)\] for $`\widehat{๐}(๐ซ,\omega )`$ the formal solution of Eq. (6.16), we derive, on using the relation (3.28),
$`\widehat{๐}_\mathrm{M}^{(+)}(๐ซ,t)=\widehat{๐}_{\mathrm{M}\mathrm{free}}^{(+)}(๐ซ,t)`$ (6.17)
$`+{\displaystyle \frac{i}{\pi \epsilon _0}}{\displaystyle _0^{\mathrm{}}}d\omega {\displaystyle \frac{\omega ^2}{c^2}}\mathrm{Im}๐ฎ(๐ซ,๐ซ_\mathrm{A},\omega )๐{\displaystyle _t^{}^t}d\tau \mathrm{e}^{i\omega (t\tau )}\widehat{\sigma }(\tau ).`$
Substitution of this expression into Eqs. (6.14) and (6.15) then yields a system of integro-differential equations for the atomic quantities.
In the Markov approximation the integro-differential equations reduce to Langevin-type differential equations. It is assumed that \[after performing the $`\omega `$-integration in Eq. (6.17)\] the time integral effectively runs over a small correlation time interval $`\tau _\mathrm{c}`$. As long as we require that $`t`$ $``$ $`t^{}`$ $``$ $`\tau _\mathrm{c}`$, we may extend the lower limit of the $`\tau `$-integral to minus infinity with little error. Further we require that $`\tau _\mathrm{c}`$ be small on a time scale on which the atomic system is changed owing to the coupling to the (medium-assisted) electromagnetic field. In this case, in the $`\tau `$-integral in Eq. (6.17) the slowly varying atomic quantity $`\widehat{\sigma }(\tau )\mathrm{e}^{i\omega _\mathrm{A}\tau }`$ can be taken at time $`t`$ and put in front of the integral, thus
$$\widehat{๐}_\mathrm{M}^{(+)}(๐ซ,t)=\widehat{๐}_{\mathrm{M}\mathrm{free}}^{(+)}(๐ซ,t)+\widehat{\sigma }(t)\frac{i}{\pi \epsilon _0}_0^{\mathrm{}}d\omega \frac{\omega ^2}{c^2}\mathrm{Im}๐ฎ(๐ซ,๐ซ_\mathrm{A},\omega )๐\zeta (\omega _\mathrm{A}\omega )$$
(6.18)
\[$`\zeta (x)`$ $`=`$ $`\pi \delta (x)`$ $`+`$ $`i๐ซx^1`$\]. Substitution of this expression into Eqs. (6.14) and (6.15) yields
$$\dot{\widehat{\sigma }}_z=\mathrm{\Gamma }(1+\widehat{\sigma }_z)+\left[\frac{2i}{\mathrm{}}\widehat{\sigma }^{}\widehat{๐}_{\mathrm{M}\mathrm{free}}^{(+)}(๐ซ_\mathrm{A})๐+\text{H.c.}\right],$$
(6.19)
$$\dot{\widehat{\sigma }}{}_{}{}^{}=[i(\omega _\mathrm{A}\delta \omega )\frac{1}{2}\mathrm{\Gamma }]\widehat{\sigma }^{}+\frac{i}{\mathrm{}}\widehat{๐}_{\mathrm{M}\mathrm{free}}^{()}(๐ซ_\mathrm{A})๐\widehat{\sigma }_z,$$
(6.20)
where
$$\mathrm{\Gamma }=\frac{2\omega _\mathrm{A}^2d_id_j}{\mathrm{}\epsilon _0c^2}\mathrm{Im}G_{ij}(๐ซ_\mathrm{A},๐ซ_\mathrm{A},\omega _\mathrm{A})$$
(6.21)
is the rate of spontaneous decay of the upper state and
$$\delta \omega =\frac{d_id_j}{\mathrm{}\pi \epsilon _0}๐ซ_0^{\mathrm{}}d\omega \frac{\omega ^2}{c^2}\frac{\mathrm{Im}G_{ij}(๐ซ_\mathrm{A},๐ซ_\mathrm{A},\omega )}{\omega \omega _\mathrm{A}}$$
(6.22)
is the contribution to the Lamb shift.<sup>22</sup><sup>22</sup>22For the the overall (vacuum) Lamb shift, see, e.g., . Note that from Eq. (3.44) it follows that Eq. (6.21) can be given in the equivalent form of
$$\mathrm{\Gamma }=\frac{2\pi }{\mathrm{}^2}d_id_j_0^{\mathrm{}}d\omega 0|\underset{ยฏ}{\overset{^}{E}}_i(๐ซ_\mathrm{A},\omega )\underset{ยฏ}{\overset{^}{E}}_j^{}(๐ซ_\mathrm{A},\omega _\mathrm{A})|0,$$
(6.23)
which exactly corresponds to Fermiโs Golden Rule (see, e.g., ).
#### 6.1.3 Schrรถdinger picture
The above given equations of motion do not only apply to the spontaneous emission but are also suitable for the study of the evolution of a two-level atom driven by an external medium-assisted electromagnetic field. If the atom is initially prepared in the upper state and there is no driving field, then the use of the wave equation
$$i\mathrm{}\frac{\mathrm{d}}{\mathrm{d}t}|\mathrm{\Psi }=\widehat{H}|\mathrm{\Psi }$$
(6.24)
may be more appropriate for the study of the motion of the coupled atomโfield system . According to the Hamiltonian in Eq. (6.2) together with Eqs. (6.4), (6.8), and (6.13), the state vector $`|\mathrm{\Psi }(t)`$ can be expanded as
$`|\mathrm{\Psi }(t)=C_u(t)\mathrm{e}^{i\omega _ut}|u|0`$ (6.25)
$`+{\displaystyle \mathrm{d}^3๐ซ_0^{\mathrm{}}d\omega C_{li}(๐ซ,\omega ,t)\mathrm{e}^{i(\omega +\omega _l)t}|l|i,๐ซ,\omega },`$
where $`|0`$ is the vacuum state of the fundamental fields $`\widehat{f}_i(๐ซ,\omega )`$, and $`|i,๐ซ,\omega `$ is the state, where one of them is excited in a single-quantum Fock state. It is not difficult to prove that the probability amplitudes $`C_u`$ and $`C_{li}`$ satisfy the differential equations
$`\dot{C}_u(t)={\displaystyle \frac{d_j}{\sqrt{\pi ฯต_0\mathrm{}}}}{\displaystyle _0^{\mathrm{}}}\mathrm{d}\omega {\displaystyle \frac{\omega ^2}{c^2}}{\displaystyle }\mathrm{d}^3๐ซ[\sqrt{ฯต_\mathrm{I}(๐ซ,\omega )}`$ (6.26)
$`\times G_{ji}(๐ซ_\mathrm{A},๐ซ,\omega )C_{li}(๐ซ,\omega ,t)\mathrm{e}^{i(\omega \omega _\mathrm{A})t}],`$
$$\dot{C}_{li}(๐ซ,\omega ,t)=\frac{d_j}{\sqrt{\pi ฯต_0\mathrm{}}}\frac{\omega ^2}{c^2}\sqrt{ฯต_\mathrm{I}(๐ซ,\omega )}G_{ji}^{}(๐ซ_\mathrm{A},๐ซ,\omega )C_u(t)\mathrm{e}^{i(\omega \omega _\mathrm{A})t}.$$
(6.27)
We now substitute the result of formal integration of Eq. (6.27) \[$`C_{li}(๐ซ,\omega ,0)`$ $`=`$ $`0`$\] into Eq. (6.26). Making use of the relationship (3.28), we obtain the integro-differential equation
$$\dot{C}_u(t)=_0^tdt^{}K(tt^{})C_u(t^{}),$$
(6.28)
with the kernel function
$$K(tt^{})=\frac{d_id_j}{\mathrm{}\pi ฯต_0}_0^{\mathrm{}}d\omega \frac{\omega ^2}{c^2}\mathrm{Im}G_{ij}(๐ซ_\mathrm{A},๐ซ_\mathrm{A},\omega )\mathrm{e}^{i(\omega \omega _\mathrm{A})(tt^{})}.$$
(6.29)
Taking the time integral of both sides of Eq. (6.28), we easily derive, on changing the order of integrations on the right-hand side,
$$C_u(t)=_0^tdt^{}\overline{K}(tt^{})C_u(t^{})+1$$
(6.30)
\[$`C_u(0)`$ $`=`$ $`1`$\], where
$$\overline{K}(tt^{})=\frac{d_id_j}{\mathrm{}\pi ฯต_0}_0^{\mathrm{}}d\omega \frac{\omega ^2}{c^2}\frac{\mathrm{Im}G_{ij}(๐ซ_\mathrm{A},๐ซ_\mathrm{A},\omega )}{i(\omega \omega _\mathrm{A})}\left[\mathrm{e}^{i(\omega \omega _\mathrm{A})(tt^{})}1\right].$$
(6.31)
Note that in the Markov approximation the kernel in Eq. (6.31) simply becomes
$$K(tt^{})=\frac{1}{2}\mathrm{\Gamma }+i\delta \omega ,$$
(6.32)
where $`\mathrm{\Gamma }`$ and $`\delta \omega `$ are respectively given by Eqs. (6.21) and (6.22).
The equation (6.30) is a well-known Volterra integral equation of the second kind. It is worth noting that the integro-differential equation (6.28) and the equivalent integral equation (6.30) apply to the spontaneous decay of an atom in the presence of an arbitrary configuration of dispersing and absorbing dielectric bodies. All the matter parameters that are relevant for the atomic evolution are contained, via the Green tensor, in the kernel functions (6.29) and (6.31). In particular when absorption is disregarded and the permittivity is regarded as being a real frequency-independent quantity (which of course can change with space), then the formalism yields the results of standard mode decomposition (see, e.g. ).
### 6.2 Atom near a dielectric surface
As a first example, let us consider the spontaneous decay of an excited atom near an absorbing planar dielectric surface. To be more specific, the distance $`z`$ to the surface of the atom is assumed to be small compared to the atomic transition wavelength. For real permittivity, this configuration has been studied extensively in connection with Casimir and van der Waals forces (see, e.g., and references cited therein). Further, it can be regarded as being the basic configuration in scanning near-field optical microscopy (see, e.g., ) and related applications.
In order to calculate the decay rate $`\mathrm{\Gamma }`$, Eq. (6.21), the imaginary part of the Green tensor of two (infinite) half-spaces with a common interface is required. When the atomโs half-space is the vacuum, then the Green tensor in this half-space can be decomposed into the vacuum Green tensor and a reflection term, i.e.,
$$G_{kl}(๐ซ,๐ซ^{},\omega )=G_{kl}^\mathrm{V}(๐ซ,๐ซ^{},\omega )+R_{kl}(๐ซ,๐ซ^{},\omega ),$$
(6.33)
where the vacuum Green tensor $`G_{kl}^\mathrm{V}(๐ซ,๐ซ^{},\omega )`$ is obtained from Eq. (A.34) for $`q`$ $`=`$ $`\omega /c`$. From Eq. (A.36) it is seen that $`\mathrm{Im}G_{kl}^\mathrm{V}(๐ซ,๐ซ^{},\omega )`$ $`=`$ $`0`$, and thus we obtain, on applying Eq. (A.38),
$$\mathrm{Im}G_{kl}^\mathrm{V}(๐ซ,๐ซ,\omega )=\frac{\omega }{6\pi c}\delta _{kl}.$$
(6.34)
Combining Eqs. (6.21), (6.33), and (6.34), we obtain
$$\mathrm{\Gamma }=\mathrm{\Gamma }_0+\frac{2\omega _\mathrm{A}^2d_kd_l}{\mathrm{}\epsilon _0c^2}\mathrm{Im}R_{kl}(๐ซ_\mathrm{A},๐ซ_\mathrm{A},\omega _\mathrm{A}),$$
(6.35)
where $`\mathrm{\Gamma }_0`$ is given by Eq. (6.1)
In the coincidence limit, the reflection term $`R_{kl}(๐ซ,๐ซ,\omega )`$ can be given in the form
$`R_{xx}(z,z,\omega )=R_{yy}(z,z,\omega )`$ (6.36)
$`={\displaystyle \frac{i}{8\pi q^2}}{\displaystyle _0^{\mathrm{}}}dkk\beta \mathrm{e}^{2i\beta z}r^p(k)+{\displaystyle \frac{i}{8\pi }}{\displaystyle _0^{\mathrm{}}}dk{\displaystyle \frac{k\mathrm{e}^{2i\beta z}}{\beta }}r^s(k),`$
$$R_{zz}(z,z,\omega )=\frac{i}{4\pi q^2}_0^{\mathrm{}}dkk^3\frac{\mathrm{e}^{2i\beta z}}{\beta }r^p(k),$$
(6.37)
where $`r^p(k)`$ and $`r^s(k)`$ are the usual Fresnel reflection coefficients for $`p`$\- and $`s`$-polarized waves ($`q`$ $`=`$ $`\omega /c`$, $`\beta `$ $`=`$ $`\sqrt{q^2k^2}`$; the origin of the co-ordinates is on the interface). When the distance $`z`$ of the atom from the surface is small compared to the wavelength, $`qz`$ $``$ $`1`$, then the integrals in Eqs.(6.36) and (6.37) can be evaluated asymptotically to obtain
$$R_{zz}=\frac{1}{16\pi q^2z^3}\frac{n^21}{n^2+1}+\frac{1}{8\pi z}\frac{(n1)^2}{n(n+1)}+\frac{iq}{12\pi }\frac{(n1)(2n1)}{n(n+1)}+O(z),$$
(6.38)
$$R_{xx}=R_{yy}=\frac{1}{2}R_{zz}\frac{1}{16\pi z}\frac{n^21}{n^2+1}\frac{iq}{3\pi }\frac{n1}{n+1}+O(z).$$
(6.39)
\[$`\epsilon `$ $`=`$ $`\epsilon (\omega )`$, $`n`$ $`=`$ $`\sqrt{\epsilon (\omega )}`$\]. Inserting Eqs. (6.38) and (6.39) into Eq. (6.35) yields ($`\omega `$ $`=`$ $`\omega _\mathrm{A}`$)
$$\mathrm{\Gamma }=\mathrm{\Gamma }_0\frac{3}{8}\left(1+\frac{d_z^2}{d^2}\right)\left(\frac{c}{\omega _\mathrm{A}z}\right)^3\frac{\epsilon _\mathrm{I}(\omega _\mathrm{A})}{|\epsilon (\omega _\mathrm{A})+1|^2}+O(z^1).$$
(6.40)
It is worth noting that the leading term $`z^3`$ in Eq. (6.40) exactly agrees with the result of the involved microscopic approach in . It typically occurs for absorbing media and reflects the possibility of nonradiative decay, if the excited atom is sufficiently near the medium.
### 6.3 Real-cavity model of spontaneous decay in a dielectric medium
Let us consider the situation when an excited atom is placed inside an empty micro-sphere (embedded in an otherwise homogeneous dielectric) with a radius much smaller than the wavelength of the atomic transition. This might serve as a model for spontaneous decay of a (guest) atom in an absorbing dielectric.<sup>23</sup><sup>23</sup>23Such a model applies as long as the atom cannot โresolveโ the atomic structure of the surrounding matter. From simple arguments based on the change of the mode density it is suggested that the spontaneous emission rate inside a nonabsorbing dielectric should be modified according to $`\mathrm{\Gamma }`$ $`=`$ $`n\mathrm{\Gamma }_0`$, where $`n`$ is the (real) refractive index of the medium and $`\mathrm{\Gamma }_0`$ is given by Eq. (6.1) . Here it is assumed that the local field the atom interacts with is exactly the same as the electromagnetic field in the continuous medium. In reality, the atom is located in a small region of free space, and the local field is different from the field in the continuous medium which leads to the introduction of the so-called local-field correction factor $`\xi `$, thus giving $`\mathrm{\Gamma }`$ $`=`$ $`n\xi \mathrm{\Gamma }_0`$. Different models have been used to compute $`\xi `$. Here we want to concentrate on the (Glauber-Lewenstein) real-cavity model, which for real $`n`$ leads to
$$\xi =\left(\frac{3n^2}{2n^2+1}\right)^2.$$
(6.41)
Experiments suggest that this model is a good candidate for describing the decay of substitutional guest atoms different from the constituents of the dielectric .
Now let us turn to the question what will appear for absorbing media. According to Eq. (6.21), the rate of spontaneous decay is proportional to the imaginary part of the Green tensor in the coincidence limit (i.e., the two spatial arguments are to be taken at the position of the atom). In the real-cavity model the Green tensor that enters the rate formula is the Green tensor for the system disturbed by a small free-space inhomogeneity. Obviously, it can again be given in a form like in Eq. (6.33), where for a sphere (with the centre as origin of co-ordinates) the reflection term $`๐น(๐ซ,๐ซ^{},\omega )`$ reads
$`๐น(๐ซ,๐ซ^{},\omega )={\displaystyle \frac{i\omega }{4\pi c}}{\displaystyle \underset{e,o}{}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m=0}{\overset{n}{}}}\{{\displaystyle \frac{2n+1}{n(n+1)}}{\displaystyle \frac{(nm)!}{(n+m)!}}(2\delta _{0m})`$ (6.42)
$`\times [C_n^M(\omega )๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,{\displaystyle \frac{\omega }{c}})๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ^{},{\displaystyle \frac{\omega }{c}})`$
$`+C_n^N(\omega )๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,{\displaystyle \frac{\omega }{c}})๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ^{},{\displaystyle \frac{\omega }{c}})]\},`$
where $`๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,k)`$ and $`๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,k)`$ are the ($`e`$ven and $`o`$dd) vector Debye potentials defined by
$$๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,k)=\mathbf{}\times \left[\psi _{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,k)๐ซ\right],$$
(6.43)
$$๐_{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,k)=\frac{1}{k}\mathbf{}\times \mathbf{}\times \left[\psi _{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,k)๐ซ\right],$$
(6.44)
with
$$\psi _{\genfrac{}{}{0pt}{}{e}{o}nm}(๐ซ,k)=j_n(kr)P_n^m(\mathrm{cos}\mathrm{\Theta })\left(\genfrac{}{}{0pt}{}{\mathrm{cos}}{\mathrm{sin}}\right)(m\varphi )$$
(6.45)
\[$`j_n(kr)`$, spherical Bessel function of the first kind; $`P_n^m(\mathrm{cos}\mathrm{\Theta })`$, associated Legendre polynomial; for the rather lengthy expressions of the generalized reflection coefficients $`C_n^{M(N)}`$, see \].
In particular in the coincidence limit $`๐ซ`$ $``$ $`๐ซ^{}`$ $``$ $`0`$ only the TM-wave vector Debye potentials $`๐_{\genfrac{}{}{0pt}{}{e}{o}10}(๐ซ,k)`$ and $`๐_{\genfrac{}{}{0pt}{}{e}{o}11}(๐ซ,k)`$ contribute to $`R_{kl}(๐ซ,๐ซ^{},\omega )`$ and we find that
$$R_{kl}(๐ซ,๐ซ,\omega )|_{๐ซ=0}=\frac{i\omega }{6\pi c}C_1^N(\omega )\delta _{kl},$$
(6.46)
where the generalized reflection coefficient $`C_1^N(\omega )`$ reads
$`C_1^N(\omega )=`$
$`{\displaystyle \frac{\left[i+\varrho (n+1)i\varrho ^2n\varrho ^3n^2/(n+1)\right]\mathrm{e}^{i\varrho }}{\mathrm{sin}\varrho \varrho (\mathrm{cos}\varrho +in\mathrm{sin}\varrho )+i\varrho ^2n\mathrm{cos}\varrho \varrho ^3(\mathrm{cos}\varrho in\mathrm{sin}\varrho )n^2/(n^21)}}`$
\[$`n`$ $`=`$ $`\sqrt{\epsilon (\omega )}`$ and $`\varrho `$ $`=`$ $`R\omega /c`$, with $`R`$ being the cavity radius\]. Hence, when the atom is located at the centre of the sphere, we may insert Eq. (6.46) into Eq. (6.35) to obtain (for $`\omega `$ $`=`$ $`\omega _\mathrm{A}`$) the decay rate in the form of<sup>24</sup><sup>24</sup>24Note that when in the coincidence limit (at the position of the atom) the Green tensor has the form of Eq. (6.33), then the rate of the spontaneous decay has the form of Eq. (6.35). Since the vacuum Green tensor $`G_{kl}^\mathrm{V}(๐ซ,๐ซ^{},\omega )`$ has no longitudinal imaginary part and the scattering component $`R_{kl}(๐ซ,๐ซ^{},\omega )`$ is purely transverse, the longitudinal electromagnetic field does not contribute to the decay rate.
$$\mathrm{\Gamma }=\mathrm{\Gamma }_0\left[1+\mathrm{Re}C_1^N(\omega _\mathrm{A})\right].$$
(6.48)
As long as the surrounding medium can be treated as a continuum this result is exact and valid for all admissible cavity sizes, without restriction to transition frequencies far from medium resonances. If the cavity radius is small compared to the wavelength of the atomic transition (but large enough for the continuum approach), we may expand the generalized reflection coefficient $`C_1^N(\omega _\mathrm{A})`$, Eq. (6.3), in powers of $`\varrho `$ $`=`$ $`R\omega _\mathrm{A}/c`$ to obtain
$`\mathrm{\Gamma }=\mathrm{\Gamma }_0\{{\displaystyle \frac{9\epsilon _\mathrm{I}(\omega _\mathrm{A})}{|2\epsilon (\omega _\mathrm{A})+1|^2}}\left({\displaystyle \frac{c}{\omega _\mathrm{A}R}}\right)^3`$ (6.49)
$`+{\displaystyle \frac{9\epsilon _\mathrm{I}(\omega _\mathrm{A})\left[28|\epsilon (\omega _\mathrm{A})|^2+16\epsilon _\mathrm{R}(\omega _\mathrm{A})+1\right]}{5|2\epsilon (\omega _\mathrm{A})+1|^4}}\left({\displaystyle \frac{c}{\omega _\mathrm{A}R}}\right)`$
$`+{\displaystyle \frac{9n_\mathrm{R}(\omega _\mathrm{A})}{|2\epsilon (\omega _\mathrm{A})+1|^4}}\left[4|\epsilon (\omega _\mathrm{A})|^4+4\epsilon _\mathrm{R}(\omega _\mathrm{A})|\epsilon (\omega _\mathrm{A})|^2+\epsilon _\mathrm{R}^2(\omega _\mathrm{A})\epsilon _\mathrm{I}^2(\omega _\mathrm{A})\right]`$
$`{\displaystyle \frac{9n_\mathrm{I}(\omega _\mathrm{A})\epsilon _\mathrm{I}(\omega _\mathrm{A})}{|2\epsilon (\omega _\mathrm{A})+1|^4}}[4|\epsilon (\omega _\mathrm{A})|^2+2\epsilon _\mathrm{R}(\omega _\mathrm{A})]\}+O({\displaystyle \frac{R\omega _\mathrm{A}}{c}}).`$
For $`\epsilon _\mathrm{I}(\omega _\mathrm{A})`$ $`=`$ $`0`$, i.e, when absorption is fully disregarded, in Eq. (6.49) only the third term survives and reproduces exactly the Glauber-Lewenstein local-field correction factor (6.41). For an absorbing medium terms proportional to $`R^3`$ and $`R^1`$ are observed, which can give rise to a strong dependence of the decay rate on the cavity radius. In particular, the leading term proportional to $`R^3`$ can be regarded as corresponding to nonradiative decay via dipole-dipole energy transfer from the atom to the medium.
Examples of the dependence of rate of spontaneous decay on the atomic transition frequency are plotted in Fig. 4 for a model permittivity of Lorentz type
$$\epsilon (\omega )=1+\frac{\omega _\mathrm{P}^2}{\omega _\mathrm{T}^2\omega ^2i\gamma \omega }.$$
(6.50)
It shows a band gap between the (transverse) resonance frequency $`\omega _\mathrm{T}`$ and the longitudinal frequency $`\omega _\mathrm{L}`$ $`=`$ $`\sqrt{\omega _\mathrm{T}^2+\omega _\mathrm{P}^2}`$. One observes, that precisely in the gap region the spontaneous decay strongly increases due to non-radiative decay channels which become important in that frequency range.
### 6.4 Cavity QED
If the radius of the microcavity is not small compared with the wavelength of the atomic transition, the cavity can act as a resonator, and it is well known that the spontaneous decay of an excited atom can be strongly modified when it is placed in a microresonator (see, e.g., ). There are typically two qualitatively different regimes: the weak coupling regime and the strong-coupling regime. In the weak coupling regime the Markov approximation applies and a monotonous exponential decay is observed, the decay rate being enhanced or reduced compared to the free-space value depending on whether the atomic transition frequency fits a cavity resonance or not. The strong-coupling regime, in contrast, is characterized by reversible Rabi oscillations where the energy of the initially excited atom is periodically exchanged between the atom and the radiation field. This usually requires that the emission is in resonance with a high-quality cavity mode. Recent progress in constructing certain types of microresonators such as microspheres has rendered it possible to approach the ultimate quality level determined by intrinsic material losses .
Let us consider the spontaneous decay of an excited atom placed inside a spherical three-layer structure (Fig. 5). The outer layer ($`r`$ $`>`$ $`R_1`$) and the inner layer ($`0`$ $``$ $`r`$ $`<`$ $`R_2`$) are vacuum, whereas the middle layer ($`R_2`$ $``$ $`r`$ $``$ $`R_1`$), which plays the role of the resonator wall, is a dispersive and absorbing dielectric. In particular when a Lorentz-type dielectric is assumed whose permittivity is of the form (6.50), the wall would be perfectly reflecting in the band-gap zone, i.e., $`\omega _\mathrm{T}`$ $`<`$ $`\omega _\mathrm{A}`$ $`<`$ $`\omega _\mathrm{L}`$, provided that absorption could be disregarded ($`\gamma `$ $``$ $`0`$). The Green tensor $`๐ฎ(๐ซ,๐ซ^{},\omega )`$ for a spherical three-layer geometry can be found in .
#### 6.4.1 Weak coupling
For $`๐ซ`$ and $`๐ซ^{}`$ inside the inner (vacuum) sphere, the Green tensor looks like the one for the two-layered spherical geometry considered in Sec. 6.3. It has again the form given in Eq. (6.33) together with Eqs. (6.42) โ (6.45), but with different reflection coefficients $`C_n^{M(N)}`$ (see ). Hence, when the atom is located at the centre of the inner sphere, then the rate of spontaneous decay can again be given in the form of Eq. (6.48). Obviously, for a sufficiently thick cavity wall, $`\mathrm{exp}[in_\mathrm{I}(\omega _\mathrm{A})(R_1R_2)\omega /c]`$ $``$ $`1`$, the generalized reflection coefficient $`C_1^N`$ in Eq. (6.48) reduces to that one for the two-layered configuration \[Eq. (6.3) with $`R`$ $`=`$ $`R_2`$\], and thus for a true microresonator, $`R_2\omega _\mathrm{A}/c`$ $``$ $`1`$, the decay rate becomes
$`\mathrm{\Gamma }\mathrm{\Gamma }_0\mathrm{Re}\left[{\displaystyle \frac{n(\omega _\mathrm{A})i\mathrm{tan}(R_2\omega _\mathrm{A}/c)}{1in(\omega _\mathrm{A})\mathrm{tan}(R_2\omega _\mathrm{A}/c)}}\right]`$ (6.51)
$`=\mathrm{\Gamma }_0{\displaystyle \frac{n_\mathrm{R}(\omega _\mathrm{A})[1+\mathrm{tan}^2(R_2\omega _\mathrm{A}/c)]}{[1+n_\mathrm{I}(\omega _\mathrm{A})\mathrm{tan}(R_2\omega _\mathrm{A}/c)]^2+n_\mathrm{R}^2(\omega _\mathrm{A})\mathrm{tan}^2(R_2\omega _\mathrm{A}/c)}}.`$
From Fig. 6 it is seen that \[for the model permittivity (6.50)\] the rate of spontaneous decay sensitively depends on the transition frequency. Narrow-band enhancement of spontaneous decay ($`\mathrm{\Gamma }/\mathrm{\Gamma }_0`$ $`>`$ $`1`$) alternates with broadband inhibition ($`\mathrm{\Gamma }/\mathrm{\Gamma }_0`$ $`<`$ $`1`$). The frequencies where the maxima of enhancement are observed correspond to the resonance frequencies of the cavity. Within the band gap the heights and widths of the frequency intervals in which spontaneous decay is feasible are essentially determined by the material losses. Outside the band-gap zone the change of the decay rate is less pronounced, because of the relatively large inputโoutput coupling, the (small) material losses being of secondary importance.
#### 6.4.2 Strong coupling
The strength of the coupling between the atom and the electromagnetic field increases when the frequency of the atomic transition frequency approaches a cavity resonance frequency. In order to gain insight into the solution of the integral equation (6.30) for the strong-coupling regime, let us consider the limiting case of only a single cavity resonance (frequency $`\omega _\mathrm{C}`$) being involved in the atomโfield interaction. In this case we may approximate, on using Eqs. (6.21) and (6.48), the kernel function (6.31) by
$`\overline{K}(tt^{}){\displaystyle \frac{\mathrm{\Gamma }_\mathrm{C}(\delta \omega _\mathrm{C})^2}{2\pi }}\mathrm{e}^{i(\omega _\mathrm{C}\omega _\mathrm{A})(tt^{})}{\displaystyle _{\mathrm{}}^{\mathrm{}}}d\omega {\displaystyle \frac{\mathrm{e}^{i(\omega \omega _\mathrm{C})(tt^{})}}{(\omega \omega _\mathrm{C})^2+\left(\delta \omega _\mathrm{C}\right)^2}}`$ (6.52)
$`=\frac{1}{2}\mathrm{\Gamma }_\mathrm{C}\delta \omega _\mathrm{C}\mathrm{e}^{i(\omega _\mathrm{C}\omega _\mathrm{A})(tt^{})}\mathrm{e}^{\delta \omega _\mathrm{C}|tt^{}|},`$
and thus the integral equation (6.30) corresponds to the differential equation
$$\ddot{C}_u(t)+\left[i(\omega _\mathrm{C}\omega _\mathrm{A})+\delta \omega _\mathrm{C}\right]\dot{C}_u(t)+\frac{1}{2}\mathrm{\Gamma }_\mathrm{C}\delta \omega _\mathrm{C}C_u(t)=0.$$
(6.53)
where $`\mathrm{\Gamma }_\mathrm{C}`$ is the decay rate at the cavity resonance \[i.e., $`\omega _\mathrm{A}`$ $`=`$ $`\omega _\mathrm{C}`$ in the rate formulas (6.48) and (6.51\], and $`\delta \omega _\mathrm{C}`$ is the width of the cavity resonance. For small absorption, $`\gamma `$ $``$ $`\omega _\mathrm{T},\omega _\mathrm{P},\omega _\mathrm{P}^2/\omega _\mathrm{T}`$, the resonance lines in the band-gap zone can be regarded as being Lorentzians, and
$$\delta \omega _\mathrm{C}=\frac{c\mathrm{\Gamma }_0}{R_2\mathrm{\Gamma }_\mathrm{C}}.$$
(6.54)
Equation (6.53) reveals that (under the assumptions made) the upper-state probability amplitude of the atom obeys the equation of motion for a damped harmonic oscillator. In the strong-coupling regime, where $`\omega _\mathrm{A}`$ $`=`$ $`\omega _\mathrm{C}`$ and $`\mathrm{\Gamma }_\mathrm{C}`$ $``$ $`\delta \omega _\mathrm{C}`$, damped Rabi oscillations are observed:<sup>25</sup><sup>25</sup>25Note that in the opposite case where $`\mathrm{\Gamma }_\mathrm{C}`$ $``$ $`\delta \omega _\mathrm{C}`$ the solution of Eq. (6.53) is $`|C_u(t)|^2`$ $`=`$ $`\mathrm{e}^{\mathrm{\Gamma }_\mathrm{C}t}`$ for $`\omega _\mathrm{A}`$ $`=`$ $`\omega _\mathrm{C}`$.
$$|C_u(t)|^2=\mathrm{e}^{\delta \omega _\mathrm{C}t}\mathrm{cos}^2(\mathrm{\Omega }t/2),$$
(6.55)
where the Rabi frequency reads
$$\mathrm{\Omega }=\sqrt{2\mathrm{\Gamma }_\mathrm{C}\delta \omega _\mathrm{C}}.$$
(6.56)
Typical examples for the time evolution of the upper-state occupation probability are shown in Fig. 7. The curves are the exact (numerical) solutions of the integral equation (6.30) \[together with the kernel function (6.31)\] for the model permittivity (6.50) The figure shows that with increasing value of the intrinsic damping constant $`\gamma `$ of the material of the cavity wall the Rabi oscillations become less pronounced. Physically, larger $`\gamma `$ means larger absorption probability of the emitted photon by the cavity wall and thus reduced probability of atom-field energy interchange.
## 7 Extensions to other media
So far we have focused our attention to quantum electrodynamics in absorbing, isotropic, non-magnetic, linear media. The quantization scheme outlined in Sec. 3 can, of course, also be generalized to other media. The starting point is the basic formula (3.35) relating the operator of the electric field strength to the dynamical variables of the system composed of the electromagnetic field and the medium. Recalling Eqs. (3.16) and (3.31), we may write Eq. (3.35) as
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=i\omega \mu _0\mathrm{d}^3๐ซ^{}๐ฎ(๐ซ,๐ซ^{},\omega )\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ^{},\omega ),$$
(7.1)
where
$$\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ,\omega )=\omega \sqrt{\frac{\mathrm{}\epsilon _0}{\pi }\epsilon _\mathrm{I}(๐ซ,\omega )}\widehat{๐}(๐ซ,\omega ).$$
(7.2)
In what follows we briefly comment on possible extensions.
### 7.1 Amplifying media
Let us consider a dielectric which in some space and frequency regions is amplifying, but in other ones absorbing. Amplification can be described by a complex permittivity $`\epsilon (๐ซ,\omega )`$ which exhibits the familiar properties as in the case of absorption, most importantly, it still fulfills the KramersโKronig relations (3.19) and (3.20), except that the imaginary part is negative, $`\epsilon _\mathrm{I}(๐ซ,\omega )`$ $`<`$ $`0`$. It can be shown that the quantization scheme in Sec. 3 also applies to amplifying media if Eq. (7.2) is modified according to
$$\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ,\omega )=\omega \sqrt{\frac{\mathrm{}\epsilon _0}{\pi }|\epsilon _\mathrm{I}(๐ซ,\omega )|}\left[\mathrm{\Theta }(\epsilon _\mathrm{I})\widehat{๐}(๐ซ,\omega )+\mathrm{\Theta }(\epsilon _\mathrm{I})\widehat{๐}^{}(๐ซ,\omega )\right]$$
(7.3)
\[$`\mathrm{\Theta }(x)`$, Heaviside step function\], which reflects the well-known fact that amplification requires the roles of the noise annihilation and creation operators to be exchanged . Substitution of this expression into Eq. (7.1) yields the representation of the ($`\omega `$-components of the) operator of the electric field strength in terms of the fundamental fields, which replaces Eq. (3.35). All the other formulas in Sec. 3 remain valid. In particular, the proof of the commutation relations (3.41) and (3.42) can be given in the same way as in Appendix B for absorbing media.
### 7.2 Anisotropic media
In anisotropic dielectrics the scalar $`\epsilon (๐ซ,\omega )`$ has to be replaced by a tensor $`๐บ(๐ซ,\omega )`$, which is symmetric for reciprocal media (see ),
$$\epsilon _{ij}(๐ซ,\omega )=\epsilon _{ji}(๐ซ,\omega ).$$
(7.4)
Thus the Green tensor has to be determined from the partial differential equation
$$\mathbf{}\times \mathbf{}\times ๐ฎ(๐ซ,๐ซ^{},\omega )\frac{\omega ^2}{c^2}๐บ(๐ซ,\omega )๐ฎ(๐ซ,๐ซ^{},\omega )=๐น(๐ซ๐ซ^{}),$$
(7.5)
which replaces Eq. (3.24). The solution for a uniaxial bulk dielectric is given in Eq. (A.39). The symmetry property (7.4) ensures that the reciprocity relation (3.27) is valid, and the generalization of Eq. (3.28) reads
$$\mathrm{d}^3๐ฌ\frac{\omega ^2}{c^2}\epsilon _{\mathrm{I}kl}(๐ฌ,\omega )G_{ik}(๐ฌ,๐ซ,\omega )G_{jl}^{}(๐ฌ,๐ซ^{},\omega )=\mathrm{Im}G_{ij}(๐ซ,๐ซ^{},\omega ).$$
(7.6)
The extension of Eq. (7.2) to anisotropic, absorbing media then reads<sup>26</sup><sup>26</sup>26Using the results of Sec. 7.1, it is not difficult to include in the scheme also amplification.
$$\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ,\omega )=\omega \sqrt{\frac{\mathrm{}\epsilon _0}{\pi }}๐บ_\mathrm{I}^{1/2}(๐ซ,\omega )\widehat{๐}(๐ซ,\omega ),$$
(7.7)
where
$$๐บ_\mathrm{I}^{1/2}(๐ซ,\omega )=๐ถ(๐ซ,\omega )๐บ_\mathrm{I}^{1/2}(๐ซ,\omega )๐ถ^1(๐ซ,\omega ).$$
(7.8)
Here the orthogonal matrix $`๐ถ(๐ซ,\omega )`$ transforms $`๐บ_\mathrm{I}(๐ซ,\omega )`$ into its (positive) diagonal form $`๐บ_\mathrm{I}^{}(๐ซ,\omega )`$,<sup>27</sup><sup>27</sup>27Note that, depending on the type of anisotropy in the dielectric, the direction of the optical axes might be frequency dependent .
$$๐บ_\mathrm{I}^{}(๐ซ,\omega )=๐ถ^1(๐ซ,\omega )๐บ_\mathrm{I}(๐ซ,\omega )๐ถ(๐ซ,\omega ).$$
(7.9)
Substitution of Eq. (7.7) into Eq. (7.1) yields the ($`\omega `$-components of the) operator of the electric field strength expressed in terms of the fundamental fields, which replaces Eq. (3.35). The further procedure is same as outlined in Sec. 3, and it can again been shown that the commutation relations (3.41) and (3.42) are fulfilled.
### 7.3 Magnetic media
Let us now include a possible magnetic response of the matter in the quantization scheme, restricting our attention to isotropic media, so that the susceptibility $`\mu (๐ซ,\omega )`$ is a scalar. We start from the constitutive relations, written in a non-standard form, which preserves the interpretation of the fields $`๐`$ and $`๐`$ as the fundamental electromagnetic fields, i.e.,
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=\epsilon _0\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega ),$$
(7.10)
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=\kappa _0\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega ).$$
(7.11)
Here, the polarization is given according to Eq. (3.9),
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=\epsilon _0\left[\epsilon (๐ซ,\omega )1\right]\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\underset{ยฏ}{\overset{^}{๐}}_\mathrm{N}(๐ซ,\omega ),$$
(7.12)
and the magnetization reads
$$\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )=\kappa _0\left[1\kappa (๐ซ,\omega )\right]\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\underset{ยฏ}{\overset{^}{๐}}_\mathrm{N}(๐ซ,\omega ),$$
(7.13)
where $`\kappa (๐ซ,\omega )`$ $`=`$ $`\mu ^1(๐ซ,\omega )`$ and $`\kappa _0`$ $`=`$ $`\mu _0^1`$. It is not difficult to prove that the Green tensor now obeys the equation
$$[\mathbf{}\times \kappa (๐ซ,\omega )\mathbf{}\times \frac{\omega ^2}{c^2}\epsilon (๐ซ,\omega )]๐ฎ(๐ซ,๐ฌ,\omega )=๐น(๐ซ๐ฌ).$$
(7.14)
Subsequently, the integral relation (3.28) has to be replaced by
$`{\displaystyle \mathrm{d}^3๐ฌ\kappa _\mathrm{I}(๐ฌ,\omega )_n^sG_{ki}(๐ฌ,๐ซ,\omega )\left[_k^sG_{nj}^{}(๐ฌ,๐ซ^{},\omega )_n^sG_{kj}^{}(๐ฌ,๐ซ^{},\omega )\right]}`$ (7.15)
$`+{\displaystyle \mathrm{d}^3๐ฌ\frac{\omega ^2}{c^2}\epsilon _\mathrm{I}(๐ฌ,\omega )G_{in}(๐ฌ,๐ซ,\omega )G_{jn}^{}(๐ฌ,๐ซ^{},\omega )}=\mathrm{Im}G_{ij}(๐ซ,๐ซ^{},\omega ),`$
which can be proved correct in analogy to the derivation given in Appendix A for non-magnetic matter.
The contribution to $`\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ,\omega )`$ of the โmagnetizationโ changes Eq. (7.2) to<sup>28</sup><sup>28</sup>28Note that Eq.(3.16) changes to $`\underset{ยฏ}{๐ฃ}_\mathrm{N}(๐ซ,\omega )`$ $`=`$ $`i\omega \underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega )`$ $`+`$ $`\mathbf{}\times \underset{ยฏ}{๐}_\mathrm{N}(๐ซ,\omega )`$.
$$\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ,\omega )=\omega \sqrt{\frac{\mathrm{}\epsilon _0}{\pi }\epsilon _\mathrm{I}(๐ซ,\omega )}\widehat{๐}_e(๐ซ,\omega )+\mathbf{}\times \sqrt{\frac{\mathrm{}\kappa _0}{\pi }|\kappa _\mathrm{I}(๐ซ,\omega )|}\widehat{๐}_m(๐ซ,\omega ),$$
(7.16)
with $`\widehat{๐}_e(๐ซ,\omega )`$ and $`\widehat{๐}_m(๐ซ,\omega )`$ being the bosonic vector fields associated with the electric and magnetic properties respectively. Here we have assumed that the medium is absorbing, so that $`\mu _\mathrm{I}(๐ซ,\omega )`$ $`>`$ $`0`$ and thus $`\kappa _\mathrm{I}(๐ซ,\omega )`$ $`=`$ $`\mu _\mathrm{I}(๐ซ,\omega )/|\mu (๐ซ,\omega )|^2`$ $`<`$ $`0`$.
Substituting for $`\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ^{},\omega )`$ in Eq. (7.1) the result of Eq. (7.16), we obtain the ($`\omega `$-components of the) operator of the electric field strength $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ expressed in terms of the dynamical variables of the system. The fields $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ and $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ can then be constructed in the same way as in Sec. 3, and by means of Eq. (7.11) the field $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ can be obtained accordingly. As before, the commutation relations (3.41) and (3.42) are fulfilled.
### 7.4 Nonlinear media
In general, the linear response of a dielectric medium to the electromagnetic field is only the first term in the series expansion of the polarization in powers of the electric field strength, and in strong fields nonlinear interactions play an important role. For example, the propagating of intense light in Kerr-type nonlinear media can lead to the formation of soliton-like pulses, the dispersion of the group velocity being compensated by the Kerr nonlinearity . There have been a number of approaches to quantize the electromagnetic field in nonlinear dielectrics , but mostly absorption is not taken into account in a consistent way.
A possible way to treat absorption is to include in the Hamiltonian (3.34) an appropriately chosen nonlinear term that may be thought of as being an integral over a nonlinear function of the medium polarization field $`๐(๐ซ)`$, and thus the extended Hamiltonian is given by
$$\widehat{H}=\mathrm{d}^3๐ซ_0^{\mathrm{}}d\omega \mathrm{}\omega \widehat{๐}^{}(๐ซ,\omega )\widehat{๐}(๐ซ,\omega )+\widehat{H}_{\mathrm{NL}}$$
(7.17)
with
$$\widehat{H}_{\mathrm{NL}}=\mathrm{d}^3๐ซh_{\mathrm{NL}}[\widehat{๐}(๐ซ)],$$
(7.18)
where
$$\widehat{๐}(๐ซ)=d\omega \underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )+\text{H.c.}$$
(7.19)
with $`\underset{ยฏ}{\overset{^}{๐}}(๐ซ,\omega )`$ being defined according to Eq. (3.9).
The Heisenberg equation of motion of $`\widehat{๐}(๐ซ,\omega )`$ then reads
$$i\mathrm{}\dot{\widehat{๐}}(๐ซ,\omega )=[\widehat{๐}(๐ซ,\omega ),\widehat{H}]=\mathrm{}\omega \widehat{๐}(๐ซ,\omega )+[\widehat{๐}(๐ซ,\omega ),\widehat{H}_{\mathrm{NL}}],$$
(7.20)
which can be rewritten as $`\widehat{\omega }\widehat{๐}(๐ซ,\omega )`$ $`=`$ $`\omega \widehat{๐}(๐ซ,\omega )`$, where
$$\widehat{\omega }=i_t+\mathrm{}^1\widehat{H}_{\mathrm{NL}}^\times ,$$
(7.21)
with $`\widehat{H}_{\mathrm{NL}}^\times \widehat{O}`$ $``$ $`[\widehat{H}_{\mathrm{NL}},\widehat{O}]`$. Hence the frequency-integrated (operator version of) Eq. (3.22) changes to
$$\mathbf{}\times \mathbf{}\times \widehat{๐}(๐ซ)\widehat{K}(๐ซ)\widehat{๐}(๐ซ)=d\omega i\omega \mu _0\underset{ยฏ}{\overset{^}{๐ฃ}}_\mathrm{N}(๐ซ,\omega )+\text{H.c.},$$
(7.22)
with $`\widehat{K}`$ being the superoperator
$$\widehat{K}(๐ซ)=\frac{1}{c^2}\left(i_t+\mathrm{}^1\widehat{H}_{\mathrm{NL}}^\times \right)^2\epsilon (๐ซ,i_t+\mathrm{}^1\widehat{H}_{\mathrm{NL}}^\times ).$$
(7.23)
A similar equation holds for the (transverse) vector potential, from which the $`\widehat{๐}`$-field can be obtained. Obviously, the approach ensures that the commutation relations (3.41) and (3.42) are satisfied.
For a linear medium, the superoperator $`\widehat{K}`$ simply describes the effects of dispersion and absorption, but for a nonlinear medium it introduces via $`\widehat{H}_{\mathrm{NL}}^\times `$ additional nonlinear coupling terms. An application of Eq. (7.22) to one-dimensional propagation of light in a Kerr-type nonlinear medium is considered in , and a simplified version has been used to study soliton propagation in absorbing optical fibres .
Acknowledgements We are grateful to Ho Trung Dung, Tomรกลก Opatrnรฝ, Eduard Schmidt, and Adriaan Tip for valuable discussions and comments. This work was supported by the Deutsche Forschungsgemeinschaft.
## A The Green tensor
### A.1 Basic relations
In Eq. (3.24), the Green tensor $`๐ฎ(๐ซ,๐ซ^{},\omega )`$ as a function of $`๐ซ`$ and $`๐ซ^{}`$ can be regarded as being the matrix elements in the position basis of a (tensor-valued) Green operator $`\widehat{๐ฎ}`$ $`=`$ $`\widehat{๐ฎ}(\omega )`$ in an abstract 1-particle Hilbert space,
$$๐ฎ(๐ซ,๐ซ^{},\omega )=๐ซ|\widehat{๐ฎ}|๐ซ^{}.$$
(A.1)
The matrix elements of the position operator $`\widehat{๐ซ}`$ are given by
$$๐ซ|\widehat{๐ซ}|๐ซ^{}=๐ซ\delta (๐ซ๐ซ^{}),$$
(A.2)
and the matrix elements of the associated momentum operator $`\widehat{๐ฉ}`$,
$$[\widehat{x}_i,\widehat{p}_j]=i\delta _{ij},$$
(A.3)
read
$$๐ซ|\widehat{๐ฉ}|๐ซ^{}=\frac{1}{i}\mathbf{}\delta (๐ซ๐ซ^{}).$$
(A.4)
Let $`\widehat{๐ฏ}`$ $`=`$ $`\widehat{๐ฏ}(\omega )`$ be the tensor-valued operator
$$\widehat{๐ฏ}=i\widehat{๐ฉ}\times i\widehat{๐ฉ}\times \widehat{q}^2\widehat{๐ฐ}=\widehat{๐ฉ}^2\widehat{๐ฐ}\widehat{๐ฉ}\widehat{๐ฉ}\widehat{q}^2\widehat{๐ฐ},$$
(A.5)
where
$$\widehat{q}^2=\frac{\omega ^2}{c^2}\epsilon (\widehat{๐ซ},\omega ),$$
(A.6)
and $`\widehat{๐ฐ}`$ is the unit operator,
$$๐ซ|\widehat{๐ฐ}|๐ซ^{}=๐น(๐ซ๐ซ^{}).$$
(A.7)
Equation (3.24) corresponds to the operator equation
$$\widehat{๐ฏ}\widehat{๐ฎ}=\widehat{๐ฐ},$$
(A.8)
as can be easily seen. Writing down Eq. (A.8) in the position basis,
$$\mathrm{d}^3๐ฌ๐ซ|\widehat{๐ฏ}|๐ฌ๐ฌ|\widehat{๐ฎ}|๐ซ^{}=๐ซ|\widehat{๐ฐ}|๐ซ^{},$$
(A.9)
and using Eqs. (A.2) and (A.4), we are just left with Eq. (3.24).
Now it is not difficult to prove Eq. (3.27). From Eq. (A.8) it follows that the equation
$$\widehat{๐ฎ}=\widehat{๐ฏ}^1$$
(A.10)
is valid, and thus we find, after multiplying it from the right by $`\widehat{๐ฏ}`$,
$$\widehat{๐ฎ}\widehat{๐ฏ}=\widehat{๐ฐ},$$
(A.11)
which in the position basis reads
$$\mathrm{d}^3๐ฌ๐ซ|\widehat{๐ฎ}|๐ฌ๐ฌ|\widehat{๐ฏ}|๐ซ^{}=๐ซ|\widehat{๐ฐ}|๐ซ^{}.$$
(A.12)
Using again Eqs. (A.2) and (A.4), after some straightforward calculation we derive
$$\left[\left(_j^r^{}_k^r^{}\delta _{jk}\mathrm{\Delta }^r^{}\right)\delta _{jk}\frac{\omega ^2}{c^2}\epsilon (๐ซ^{},\omega )\right]G_{ik}(๐ซ,๐ซ^{},\omega )=\delta _{ji}\delta (๐ซ^{}๐ซ).$$
(A.13)
Comparing Eq. (A.13) with Eq. (3.25), we immediately see the Green tensor indeed obeys Eq. (3.27).
In order to prove Eq. (3.28), we introduce operators $`\widehat{๐ถ}^{}`$ defined by
$$(\widehat{O}{}_{}{}^{})_{ij}=\left(\widehat{O}_{ji}\right){}_{}{}^{}\widehat{O}_{ji}^{}.$$
(A.14)
From Eq. (A.8) it then follows that
$$\widehat{๐ฎ}{}_{}{}^{}\widehat{๐ฏ}{}_{}{}^{}=\widehat{๐ฐ}.$$
(A.15)
Multiplying Eq. (A.8) from the left by $`\widehat{๐ฎ}^{}`$ and Eq. (A.15) from the right by $`\widehat{๐ฎ}`$ and subtracting the resulting equations from each other, we find
$$\widehat{๐ฎ}{}_{}{}^{}(\widehat{๐ฏ}\widehat{๐ฏ}{}_{}{}^{})\widehat{๐ฎ}=\widehat{๐ฎ}{}_{}{}^{}\widehat{๐ฎ},$$
(A.16)
which in Cartesian components reads
$$\widehat{G}_{mi}^{}\left(\widehat{H}_{mn}\widehat{H}_{nm}^{}\right)\widehat{G}_{nj}=\widehat{G}_{ji}^{}\widehat{G}_{ij}.$$
(A.17)
From Eq. (A.5) it is easily seen that
$$\widehat{H}_{mn}\widehat{H}_{nm}^{}=\delta _{mn}\left(\widehat{q}^2\widehat{q}^2\right).$$
(A.18)
Representing Eq. (A.17) \[together with Eq. (A.18)\] in the position basis and recalling the relations
$$๐ซ|\widehat{G}_{ji}^{}|๐ซ^{}=๐ซ^{}|\widehat{G}_{ji}|๐ซ^{}=G_{ji}^{}(๐ซ^{},๐ซ,\omega )=G_{ij}^{}(๐ซ,๐ซ^{},\omega ),$$
(A.19)
we eventually arrive at Eq. (3.28).
#### A.1.1 Asymptotic behaviour
The Green tensor $`G_{ij}(๐ซ,๐ซ^{},\omega )`$ is holomorphic in the upper half-plane of complex $`\omega `$, because of the holomorphic behaviour of $`\epsilon (๐ซ,\omega )`$. In order to study the behaviour of $`G_{ij}(๐ซ,๐ซ^{},\omega )`$ for $`|\omega |`$ $``$ $`\mathrm{}`$ and $`|\omega |`$ $``$ $`0`$, we first introduce the tensor-valued projection operators
$$\widehat{๐ฐ}{}_{}{}^{}=\widehat{๐ฐ}\widehat{๐ฐ}{}_{}{}^{},\widehat{๐ฐ}{}_{}{}^{}=\frac{\widehat{๐ฉ}\widehat{๐ฉ}}{\widehat{๐ฉ}^2}$$
(A.20)
and decompose $`\widehat{๐ฏ}`$, Eq. (A.5), as
$$\widehat{๐ฏ}=(\widehat{๐ฉ}{}_{}{}^{2}\widehat{q}{}_{}{}^{2})\widehat{๐ฐ}{}_{}{}^{}\widehat{q}{}_{}{}^{2}\widehat{๐ฐ}{}_{}{}^{}.$$
(A.21)
Applying the Feshbach formula , we then may decompose the Green tensor operator $`\widehat{๐ฎ}`$ $`=`$ $`\widehat{๐ฏ}^1`$, Eq. (A.10), as
$`\widehat{๐ฎ}=\widehat{๐ฏ}{}_{}{}^{1}=\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}^{}`$ (A.22)
$`+[\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}){}_{}{}^{1}\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}]\widehat{๐ฒ}[\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}){}_{}{}^{1}\widehat{๐ฐ}{}_{}{}^{}],`$
where
$$\widehat{๐ฒ}=[\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}){}_{}{}^{1}\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}]^1.$$
(A.23)
It is not difficult to prove that
$$\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}=\widehat{๐ฐ}{}_{}{}^{}\widehat{q}{}_{}{}^{2}\widehat{๐ฐ}{}_{}{}^{},\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}=\widehat{๐ฐ}{}_{}{}^{}\widehat{q}{}_{}{}^{2}\widehat{๐ฐ}{}_{}{}^{},$$
(A.24)
$$\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}=\widehat{๐ฐ}{}_{}{}^{}\widehat{q}{}_{}{}^{2}\widehat{๐ฐ}{}_{}{}^{},\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฏ}\widehat{๐ฐ}{}_{}{}^{}=\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฉ}{}_{}{}^{2}\widehat{q}{}_{}{}^{2})\widehat{๐ฐ}{}_{}{}^{}.$$
(A.25)
Combining Eqs. (A.22) โ (A.25) and using Eq. (A.6), we obtain for $`\widehat{๐ฎ}`$ the expression
$`\widehat{๐ฎ}={\displaystyle \frac{c^2}{\omega ^2}}\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}^{}`$ (A.26)
$`+[\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{}]\widehat{๐ฒ}[\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}{}_{}{}^{}],`$
where
$$\widehat{๐ฒ}=[\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฉ}^2\frac{\omega ^2}{c^2}\widehat{\epsilon })\widehat{๐ฐ}{}_{}{}^{}+\frac{\omega ^2}{c^2}\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }\widehat{๐ฐ}{}_{}{}^{}]^1$$
(A.27)
\[$`\widehat{\epsilon }`$ $`=`$ $`\epsilon (\widehat{๐ซ},\omega )`$\].
Now the desired limiting processes can be performed easily. For $`|\omega |`$ $``$ $`0`$ we find \[$`\widehat{\epsilon }^{(0)}`$ $`=`$ $`\epsilon ^{(0)}(\widehat{๐ซ})`$ $`=`$ $`\epsilon (\widehat{๐ซ},\omega =0)`$\]
$$\underset{|\omega |0}{lim}\frac{\omega ^2}{c^2}\widehat{๐ฎ}=\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }_{}^{(0)}\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}^{}$$
(A.28)
and
$$\underset{|\omega |0}{lim}\widehat{q}{}_{}{}^{2}\widehat{๐ฎ}=\widehat{\epsilon }^{(0)}\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }_{}^{(0)}\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}{}_{}{}^{}.$$
(A.29)
For $`|\omega |`$ $``$ $`\mathrm{}`$ we arrive at, on recalling that $`\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}^{}`$ $`=`$ $`\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}^{}`$ $`=`$ $`0`$ and $`\epsilon (๐ซ,\omega )`$ $``$ $`1`$ if $`|\omega |`$ $``$ $`\mathrm{}`$,
$$\underset{|\omega |\mathrm{}}{lim}\frac{\omega ^2}{c^2}\widehat{๐ฎ}=\underset{|\omega |\mathrm{}}{lim}\widehat{q}{}_{}{}^{2}\widehat{๐ฎ}=\widehat{๐ฐ}.$$
(A.30)
Obviously, the first term on the right hand of Eq. (A.26) is the singular part of $`\widehat{๐ฎ}`$ for $`|\omega |`$ $``$ $`0`$. Performing in that term the Taylor expansion
$$\epsilon (\widehat{๐ซ},\omega )=\epsilon ^{(0)}(\widehat{๐ซ})+\omega \epsilon ^{(1)}(\widehat{๐ซ})+\mathrm{},$$
(A.31)
where $`\epsilon ^{(0)}(\widehat{๐ซ})`$ is real and $`\epsilon ^{(1)}(\widehat{๐ซ})`$ is imaginary \[see Eq. (3.21)\], we find that
$$\mathrm{Re}G_{ij}(๐ซ,๐ซ^{},\omega )\omega ^2(|\omega |0)$$
(A.32)
and
$$\mathrm{Im}G_{ij}(๐ซ,๐ซ^{},\omega )\omega ^1(|\omega |0).$$
(A.33)
#### A.1.2 Isotropic bulk material
For a bulk material of given permittivity $`\epsilon (\omega )`$ the Green tensor reads ($`๐`$ $`=`$ $`๐ซ`$ $``$ $`๐ซ^{}`$)
$$๐ฎ(๐ซ,๐ซ^{},\omega )=\left[\mathbf{}^r\mathbf{}^r+๐ฐq^2(\omega )\right]\frac{\mathrm{e}^{iq(\omega )\rho }}{4\pi q^2(\omega )\rho },$$
(A.34)
where
$$q(\omega )=\sqrt{\epsilon (\omega )}\frac{\omega }{c}=\left[n_\mathrm{R}(\omega )+in_\mathrm{I}(\omega )\right]\frac{\omega }{c}.$$
(A.35)
It can be split up into a longitudinal and a transverse part according to
$$๐ฎ^{}(๐ซ,๐ซ^{},\omega )=\frac{1}{4\pi q^2}\left[\frac{4\pi }{3}\delta (๐)๐ฐ+\left(๐ฐ\frac{3๐๐}{\rho ^2}\right)\frac{1}{\rho ^3}\right]$$
(A.36)
and
$`๐ฎ^{}(๐ซ,๐ซ^{},\omega )={\displaystyle \frac{1}{4\pi q^2}}\{(๐ฐ{\displaystyle \frac{3๐๐}{\rho ^2}}){\displaystyle \frac{1}{\rho ^3}}`$ (A.37)
$`+q^3[({\displaystyle \frac{1}{q\rho }}+{\displaystyle \frac{i}{(q\rho )^2}}{\displaystyle \frac{1}{(q\rho )^3}})๐ฐ({\displaystyle \frac{1}{q\rho }}+{\displaystyle \frac{3i}{(q\rho )^2}}{\displaystyle \frac{3}{(q\rho )^3}}){\displaystyle \frac{๐๐}{\rho ^2}}]\mathrm{e}^{iq\rho }\}.`$
In particular, from Eq. (A.37) it follows that
$$\mathrm{Im}๐ฎ^{}(๐ซ,๐ซ,\omega )=\underset{๐ซ^{}๐ซ}{lim}\mathrm{Im}๐ฎ^{}(๐ซ,๐ซ^{},\omega )=\frac{\omega }{6\pi c}n_\mathrm{R}(\omega )๐ฐ.$$
(A.38)
#### A.1.3 Uniaxial bulk material
Following , the Green tensor for a homogeneous uniaxial dielectric can be given in the form of
$`๐ฎ(๐ซ,๐ซ^{},\omega )=q_t^2(\omega )\left[\mathbf{}^r\mathbf{}^r+q_t^2(\omega )\epsilon _c๐บ^1\right]{\displaystyle \frac{\mathrm{e}^{iq_t(\omega )\rho _e}}{4\pi \rho _e}}`$ (A.39)
$`\left[{\displaystyle \frac{\epsilon _c}{\epsilon _t}}{\displaystyle \frac{\mathrm{e}^{iq_t(\omega )\rho _e}}{4\pi \rho _e}}{\displaystyle \frac{\mathrm{e}^{iq_t(\omega )\rho }}{4\pi \rho }}\right]{\displaystyle \frac{(๐\times ๐)(๐\times ๐)}{(๐\times ๐)^2}}`$
$`\left[{\displaystyle \frac{\mathrm{e}^{iq_t(\omega )\rho _e}\mathrm{e}^{iq_t(\omega )\rho }}{4\pi iq_t(\omega )}}\right]\left[{\displaystyle \frac{๐ฐ๐๐}{(๐\times ๐)^2}}{\displaystyle \frac{2(๐\times ๐)(๐\times ๐)}{(๐\times ๐)^4}}\right],`$
\[$`\rho _e^2`$ $`=`$ $`\epsilon _c๐๐บ^1๐`$, $`q_t`$ $`=`$ $`\sqrt{\epsilon _t}\omega /c`$\]. Here, the vector $`๐`$ is the unit vector parallel to the crystallographic axis of the medium, $`\epsilon _c`$ $`=`$ $`\epsilon _c(\omega )`$ the (complex) permittivity in that direction, and $`\epsilon _t`$ $`=`$ $`\epsilon _t(\omega )`$ the (complex) permittivity transverse to it. Note that for $`\epsilon _c`$ $`=`$ $`\epsilon _t`$ $`=`$ $`\epsilon `$ \[and $`๐บ^1`$ $`=`$ $`\epsilon ^1๐ฐ`$\] the Green tensor (A.34) for isotropic bulk material is recovered.
## B Commutation relations
Let us first consider the commutation relations of the electric field and the vector potential. Using Eqs. (3.38) and (3.46) together with Eqs. (3.35) and (3.47), recalling the commutation relations (3.32) and (3.33), and applying the integral relation (3.28), after some algebra we derive
$$[\widehat{E}_k(๐ซ),\widehat{E}_l(๐ซ^{})]=0,$$
(B.1)
$$[\widehat{A}_k(๐ซ),\widehat{A}_l(๐ซ^{})]=0,$$
(B.2)
$$[\epsilon _0\widehat{E}_k(๐ซ),\widehat{A}_l(๐ซ^{})]=\mathrm{d}^3๐ฌ\left[\frac{2i\mathrm{}}{\pi }_0^{\mathrm{}}d\omega \frac{\omega }{c^2}\mathrm{Im}G_{km}(๐ซ,๐ฌ,\omega )\right]\delta _{ml}^{}(๐ฌ๐ซ^{}).$$
(B.3)
In Eq. (B.3) the $`\omega `$-integral can be performed by applying the rule
$$_0^{\mathrm{}}d\omega \mathrm{}=\underset{ฯต0}{lim}_ฯต^{\mathrm{}}d\omega \mathrm{}.$$
(B.4)
Note that, according to Eq. (A.33), $`\mathrm{Im}G_{il}(๐ซ,๐ซ^{},\omega )`$ behaves like $`\omega ^1`$ as $`\omega `$ approaches zero. Thus we may transform, on recalling Eq. (3.26), the $`\omega `$-integral into a principal-part ($`๐ซ`$) integral, so that Eq.(B.3) reads
$$[\epsilon _0\widehat{E}_k(๐ซ),\widehat{A}_l(๐ซ^{})]=\mathrm{d}^3๐ฌ\left[\mathrm{}\frac{๐ซ}{\pi }_{\mathrm{}}^{\mathrm{}}\frac{\mathrm{d}\omega }{\omega }\frac{\omega ^2}{c^2}G_{km}(๐ซ,๐ฌ,\omega )\right]\delta _{ml}^{}(๐ฌ๐ซ^{}).$$
(B.5)
The evaluation of the $`\omega `$-integral in Eq. (B.5) can be performed by means of contour-integral techniques. Since $`G_{km}(๐ซ,๐ฌ,\omega )`$ is a holomorphic function of $`\omega `$ in the upper complex frequency half-plane with the asymptotic behaviour according to Eq. (A.30), the $`\omega `$-integrals can be calculated by contour integration along an infinitely small half-circle $`|\omega |`$ $`=`$ $`\rho `$, $`\rho `$ $``$ $`0`$, and an infinitely large half-circle $`|\omega |`$ $`=`$ $`R`$, $`R`$ $``$ $`\mathrm{}`$, in the upper complex half-plane
$$๐ซ_{\mathrm{}}^{\mathrm{}}d\omega \mathrm{}=\underset{\rho 0}{lim}_{\genfrac{}{}{0pt}{}{|\omega |=\rho }{\omega _\mathrm{I}>0}}d\omega \mathrm{}\underset{R\mathrm{}}{lim}_{\genfrac{}{}{0pt}{}{|\omega |=R}{\omega _\mathrm{I}>0}}d\omega \mathrm{}.$$
(B.6)
The definition of $`\widehat{๐ฐ}^{()}`$ as given in Eq. (A.20) implies that
$$๐ซ|\widehat{๐ฐ}{}_{}{}^{()}|๐ซ^{}=๐น^{()}(๐ซ๐ซ^{}),$$
(B.7)
and thus we may write, on applying Eq. (A.1),
$$\mathrm{d}^3๐ฌ\left[_๐\frac{\mathrm{d}\omega }{\omega }\frac{\omega ^2}{c^2}G_{km}(๐ซ,๐ฌ,\omega )\right]\delta _{ml}^{}(๐ฌ๐ซ^{})=_๐\frac{\mathrm{d}\omega }{\omega }\frac{\omega ^2}{c^2}๐ซ|\widehat{๐ฎ}(\omega )\widehat{๐ฐ}{}_{}{}^{}|๐ซ^{}.$$
(B.8)
Note that
$$๐น^{}(๐ซ)=\mathbf{}\mathbf{}\frac{1}{4\pi |๐ซ|}$$
(B.9)
and
$$๐น^{}(๐ซ)=๐น(๐ซ)๐น^{}(๐ซ).$$
(B.10)
From Eq. (A.28) it follows that
$$\underset{|\omega |0}{lim}\frac{\omega ^2}{c^2}\widehat{๐ฎ}\widehat{๐ฐ}{}_{}{}^{}=\widehat{๐ฐ}{}_{}{}^{}(\widehat{๐ฐ}{}_{}{}^{}\widehat{\epsilon }_{}^{(0)}\widehat{๐ฐ}{}_{}{}^{})_{}^{1}\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}{}_{}{}^{}=0$$
(B.11)
($`\widehat{๐ฐ}{}_{}{}^{}\widehat{๐ฐ}^{}`$ $`=`$ $`0`$) and therefore the integral over the small half-circle vanishes. Finally, from Eq. (A.30) we see that
$$\underset{|\omega |\mathrm{}}{lim}\frac{\omega ^2}{c^2}\widehat{๐ฎ}\widehat{๐ฐ}{}_{}{}^{}=\widehat{๐ฐ}{}_{}{}^{}.$$
(B.12)
Hence,
$`๐ซ{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{\mathrm{d}\omega }{\omega }}{\displaystyle \frac{\omega ^2}{c^2}}๐ซ|\widehat{๐ฎ}(\omega )\widehat{๐ฐ}{}_{}{}^{}|๐ซ^{}`$ (B.13)
$`=\mathrm{\hspace{0.17em}2}i{\displaystyle _0^{\mathrm{}}}\mathrm{d}\omega {\displaystyle \frac{\omega }{c^2}}\mathrm{Im}๐ซ|\widehat{๐ฎ}(\omega )\widehat{๐ฐ}{}_{}{}^{}|๐ซ^{}=i\pi ๐น^{}(๐ซ๐ซ^{}),`$
and the sought commutator reads
$$[\epsilon _0\widehat{E}_k(๐ซ),\widehat{A}_l(๐ซ^{})]=i\mathrm{}\delta _{kl}^{}(๐ซ๐ซ^{}).$$
(B.14)
Since the corresponding commutation relations for the displacement field \[Eq. (3.40) together with Eqs. (3.37) and (3.35)\] can derived in a quite similar way, we renounce the derivation here and immediately give the result
$$[\widehat{D}_k(๐ซ),\widehat{D}_l(๐ซ^{})]=0,$$
(B.15)
$$[\widehat{D}_k(๐ซ),\widehat{A}_l(๐ซ^{})]=i\mathrm{}\delta _{kl}^{}(๐ซ๐ซ^{}).$$
(B.16)
This is the result we have expected, because the polarization $`\widehat{๐}(๐ซ)`$ $`=`$ $`\widehat{๐}(๐ซ)`$ $``$ $`\epsilon _0\widehat{๐}(๐ซ)`$ is related to the degrees of freedom of the matter and it should therefore commute with the radiation field operators.
Recalling the relations $`\mathbf{}\times \widehat{๐}`$ $`=`$ $`\widehat{๐}`$, $`\widehat{๐ท}`$ $`=`$ $`\epsilon _0\widehat{๐}^{}`$, $`\mathbf{}\widehat{\phi }`$ $`=`$ $`\widehat{๐}^{}`$, and $`\widehat{๐}`$ $`=`$ $`\widehat{๐}`$ $``$ $`\epsilon _0\widehat{๐}`$ and using the commutation relations (B.1), (B.2), (B.14), (B.15), and (B.16), we can then derive further commutation relations such as
$$[\epsilon _0\widehat{E}_k(๐ซ),\widehat{B}_l(๐ซ^{})]=i\mathrm{}ฯต_{klm}_m^r\delta (๐ซ๐ซ^{}),$$
(B.17)
$$[\widehat{A}_k(๐ซ),\widehat{\mathrm{\Pi }}_l(๐ซ^{})]=i\mathrm{}\delta _{kl}^{}(๐ซ๐ซ^{}),$$
(B.18)
$$[\widehat{\phi }(๐ซ),\widehat{\phi }(๐ซ^{})]=[\widehat{\phi }(๐ซ),\widehat{A}_k(๐ซ^{})]=[\widehat{\phi }(๐ซ),\widehat{E}_k(๐ซ^{})]=0,$$
(B.19)
$$[\widehat{\phi }(๐ซ),\widehat{D}_k(๐ซ^{})]=0.$$
(B.20)
It should be pointed out that the commutation relations given above are valid for both the medium-assisted electromagnetic field considered in Sec. 3 and the total electromagnetic field considered in Sec. 4. Needless to say that quantities of the medium-assisted electromagnetic field and quantities of the additional charged particles commute.
## C Equations of motion
In order to prove that the quantization scheme yields the correct Maxwell equations and the Newtonian equations of motion of the charged particles, let us consider, e.g. the Maxwell equation (4.8). Using the minimal-coupling Hamiltonian (4.1), we find that
$`\dot{\widehat{๐}}(๐ซ)={\displaystyle \frac{1}{i\mathrm{}}}[\widehat{๐}(๐ซ),\widehat{H}]={\displaystyle \frac{1}{i\mathrm{}}}{\displaystyle \mathrm{d}^3๐ซ^{}_0^{\mathrm{}}d\omega \mathrm{}\omega [\widehat{๐}(๐ซ),\widehat{๐}^{}(๐ซ^{},\omega )\widehat{๐}(๐ซ^{},\omega )]}`$ (C.1)
$`+{\displaystyle \frac{1}{i\mathrm{}}}{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}[\widehat{๐}(๐ซ),\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right]^2]`$
$`+{\displaystyle \frac{1}{2i\mathrm{}}}{\displaystyle \mathrm{d}^3๐ซ^{}[\widehat{๐}(๐ซ),\widehat{\rho }_\mathrm{A}(๐ซ^{})\widehat{\phi }_\mathrm{A}(๐ซ^{})]}+{\displaystyle \frac{1}{i\mathrm{}}}{\displaystyle \mathrm{d}^3๐ซ^{}[\widehat{๐}(๐ซ),\widehat{\rho }_\mathrm{A}(๐ซ^{})\widehat{\phi }_\mathrm{M}(๐ซ^{})]}.`$
Recalling the definition (4.13) of the displacement field and applying the commutation relation (B.20), we may simplify Eq. (C.1) to
$`\dot{\widehat{๐}}(๐ซ)={\displaystyle \frac{1}{i\mathrm{}}}{\displaystyle \mathrm{d}^3๐ซ^{}_0^{\mathrm{}}d\omega \mathrm{}\omega [\widehat{๐}_\mathrm{M}(๐ซ),\widehat{๐}^{}(๐ซ^{},\omega )\widehat{๐}(๐ซ^{},\omega )]}`$ (C.2)
$`+{\displaystyle \frac{1}{i\mathrm{}}}{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}[\widehat{๐}_\mathrm{M}(๐ซ),\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right]^2]`$
$`{\displaystyle \frac{\epsilon _0}{i\mathrm{}}}{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}[\mathbf{}\widehat{\phi }_\mathrm{A}(๐ซ),\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right]^2].`$
Using Eqs. (3.39) and (3.40) together with Eqs. (3.36), (3.37), and (3.35) and recalling the basic commutation relations (3.32) and (3.33), we find that the first of the three commutators in Eq. (C.2) gives
$$\frac{1}{i\mathrm{}}\mathrm{d}^3๐ซ^{}_0^{\mathrm{}}d\omega \mathrm{}\omega [\widehat{๐}_\mathrm{M}(๐ซ),\widehat{๐}^{}(๐ซ^{},\omega )\widehat{๐}(๐ซ^{},\omega )]=\frac{1}{\mu _0}\mathbf{}\times \widehat{๐}(๐ซ).$$
(C.3)
In order to calculate the second commutator, we use the commutation relation (B.16) and obtain
$$\frac{1}{i\mathrm{}}\underset{\alpha }{}\frac{1}{2m_\alpha }[\widehat{๐}_\mathrm{M}(๐ซ),\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right]^2]=\widehat{๐ฃ}_\mathrm{A}^{}(๐ซ).$$
(C.4)
Finally, recalling the definitions of $`\widehat{\phi }_\mathrm{A}(๐ซ)`$ \[Eq. (4.2)\] and $`\widehat{\rho }_\mathrm{A}(๐ซ)`$ \[Eq. (4.3)\] and using the particle-operator commutation relations, we derive
$$\frac{\epsilon _0}{i\mathrm{}}\underset{\alpha }{}\frac{1}{2m_\alpha }[\mathbf{}\widehat{\phi }_\mathrm{A}(๐ซ),\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right]^2]=\widehat{๐ฃ}_\mathrm{A}^{}(๐ซ).$$
(C.5)
In Eqs. (C.4) and (C.5), the transverse and longitudinal current densities $`\widehat{๐ฃ}_\mathrm{A}^{}(๐ซ)`$ and $`\widehat{๐ฃ}_\mathrm{A}^{}(๐ซ)`$, respectively, are defined by
$$\widehat{๐ฃ}_\mathrm{A}^{()}(๐ซ)=\frac{1}{2}\underset{\alpha }{}q_\alpha \dot{\widehat{๐ซ}}_\alpha ๐น^{()}(๐ซ\widehat{๐ซ}_\alpha )+\text{H.c.},$$
(C.6)
where the velocity operator $`\dot{\widehat{๐ซ}}_\alpha `$ of the $`\alpha `$th particle is
$$\dot{\widehat{๐ซ}}_\alpha =\frac{1}{i\mathrm{}}[\widehat{๐ซ}_\alpha ,\widehat{H}]=\frac{1}{m_\alpha }\left[\widehat{๐ฉ}_\alpha q_\alpha \widehat{๐}(\widehat{๐ซ}_\alpha )\right].$$
(C.7)
Substituting Eqs. (C.3), (C.4), and (C.5) back into Eq. (C.2), we just arrive at the Maxwell equation (4.8). The Maxwell equation (4.6) and the Newtonian equations of motion (4.10) and (4.11) can be derived in a quite similar way.
|
warning/0006/cond-mat0006020.html
|
ar5iv
|
text
|
# Magnetization plateaus of the Shastry-Sutherland model for SrCu2(BO3)2: Spin-density wave, supersolid and bound states
## I Introduction
Since plateau structures were observed in the magnetization process of a series of quasi one-dimensional Ni-compounds, magnetization plateaus have been attracting extensive interests. The appearance of plateaus in magnetization curves was explained as metal-insulator transitions of magnetic excitations driven by a magnetic field; Magnetic excitations crystallize and form SDW in the plateau states, and they are itinerant in the non-plateau states. Recently it was discussed that this phenomenon is not limited to the one-dimensional systems, but more general, and occurs in two- and three-dimensional systems as well. Before the recent studies, magnetization plateaus were already known to appear at $`m/m_{\mathrm{sat}}=1/3`$ in the antiferromagnetic compounds on the triangular lattice, C<sub>6</sub>Eu, CsCuCl<sub>3</sub>, and RbFe(MoO<sub>4</sub>)<sub>2</sub>. Theoretically, plateaus were seen at $`m/m_{\mathrm{sat}}=1/3`$ in the Heisenberg antiferromagnet on the triangular lattice, and also at $`m/m_{\mathrm{sat}}=1/2`$ in the multiple-spin exchange model with four-spin interactions. The 1/3-plateau comes from appearance of a collinear uud state and the 1/2-plateau from the uuud state. These magnetization plateaus can also be regarded as superfluid-insulator transitions of flipped-spin degree of freedom. It may be worth mentioning that there is also another trial to realize magnetization plateau in two-dimensional systems as gapped spin liquid states analogous to the FQHE wave functions.
Recently a quasi two-dimensional compound SrCu<sub>2</sub>(BO<sub>3</sub>)<sub>2</sub> is attracting extensive interests because it shows magnetization plateaus and peculiar dynamical properties. The two-dimensional lattice structure of Cu<sup>2+</sup> ions in SrCu<sub>2</sub>(BO<sub>3</sub>)<sub>2</sub> is the so-called Shastry-Sutherland lattice, which is shown in Fig. 1. Susceptibility and specific heat data show that interactions are antiferromagnetic, the spin excitation has a gap above the ground state, and the spin anisotropy is weak. This material seems to be well described with the $`S=1/2`$ Heisenberg antiferromagnet on the Shastry-Sutherland lattice. (Hereafter we call this model simply as Shastry-Sutherland model.) The ground state of the Shastry-Sutherland model is exactly a direct product of local dimer singlets on bonds $`J`$ for the region $`J^{}/J<0.68`$ (Refs. ) and there is a finite gap above the ground state. Susceptibility and specific heat estimated from this model with $`J^{}/J=0.68`$ fit well with the experimental results. (Recently the value has been modified to $`J^{}/J=0.635`$ by taking into account the three-dimensional coupling of Shastry-Sutherland layers.) In ref. , Kageyama et al. reported two plateaus at $`m/m_{\mathrm{sat}}=1/8`$ and $`1/4`$ in the magnetization curve of SrCu<sub>2</sub>(BO<sub>3</sub>)<sub>2</sub>. Theoretically, we studied the magnetization process of the Shastry-Sutherland model, treating a dimer triplet as a particle, and thereby predicted a novel broad plateau at $`m/m_{\mathrm{sat}}=1/3`$ in the previous report. It was argued that the appearance is due to superfluid-insulator transition of the excitations. Quite recently, the above 1/3-plateau was experimentally observed in magnetization measurements up to a strong field of 57\[T\]. As was predicted in ref. , this plateau is the broadest one ever found in this material. This seems to support the correctness of our argument based on the particle picture. This material also shows peculiar dynamical properties, e.g. one-magnon excitation is almost dispersionless, but two-magnon excitations have strong dispersion. The aim of this paper is to present the details of our analyses and results reported briefly in ref. , and to proceed further thereby giving remarkable consequences of the correlating hopping of the effective Hamiltonian. This correlated hopping can also explain the peculiar dynamical behaviors observed in experiments.
The ground state of the Shastry-Sutherland model at zero magnetic field was studied for a varying ratio $`J^{}/J`$ by the mean field approximation, exact diagonalization method, and series expansion. It was found that the ground state is the exact dimer state for $`J^{}/J<0.69`$, a gapped plaquette singlet state for $`0.69<J^{}/J<0.86`$, and Nรฉel-ordered state for $`0.86<J^{}/J`$. The parameters of SrCu<sub>2</sub>(BO<sub>3</sub>)<sub>2</sub> estimated as $`J^{}/J=0.68`$ (or 0.635) suggest that the spin state of the real material belongs to the dimer phase, but it is very close to the phase boundary with the plaquette singlet phase. Consistency between theoretical and experimental results on magnetization plateau at $`m/m_{\mathrm{sat}}=1/3`$ and dynamical behavior in inelastic neutron scattering also supports that the real material is in the dimer phase.
In this paper, we study the $`S=1/2`$ Heisenberg antiferromagnet on the Shastry-Sutherland lattice and discuss the magnetic properties under a magnetic field. We analyze this model using strong-coupling expansion. In section II, we derive an effective Hamiltonian for the dimer-triplet excitations. Virtual triplet excitations yield various effective repulsive interactions, which are responsible for the plateaus. Although the usual single-particle hopping is completely missing from the resulting Hamiltonian, correlated hopping processes are contained instead.
In section III, we investigate the magnetization process applying the classical approximation to the effective (pseudo-spin) Hamiltonian and show that plateaus appear at $`m/m_{\mathrm{sat}}=1/3`$ and $`1/2`$. Both of the plateau states are Mott (SDW) insulators of magnetic excitations. Spatially anisotropic interactions perturbatively generated stabilize several SDW structures. Near the plateau states, there are supersolid phases, in which SDW long-range order (LRO) and superfluid coexist. Field ($`B`$) versus coupling ($`J^{}/J`$) phase diagram is also presented.
Although the inter-dimer coupling $`J^{}/J`$ is not small, the special geometry of the Shastry-Sutherland lattice strongly suppresses bare one-particle hopping and hence the correlated hopping becomes important. This fact leads to remarkable consequences on the motion of magnetic excitations. In section IV, we investigate correlated hopping more closely and demonstrate how it favors the formation of a bound pair of dimer triplets which repel each other at the level of the bare Hamiltonian. This bound state has a relatively large dispersion and may be an elementary particle at very low magnetization, instead of the single dimer triplet. The non-plateau state would be superfluid of these bound states at least for very small magnetization. Above a certain threshold value of magnetization, individual dimer triplets become elementary particles and the non-plateau state is characterized by superfluidity of single dimer triplets.
In section V, we also discuss spin excitation just below the saturation field. It is found that lowest energy states exist on a close curve instead of a point in the momentum space.
Critical phenomena of the plateau transition are discussed in section VI and are argued to be in the same universality class as the superfluid transition of the interacting boson system in the dilute limit.
## II Effective Hamiltonian
In this section we derive an effective Hamiltonian for the magnetic excitations under a strong enough magnetic field. We begin with the $`J^{}=0`$ limit. In this limit, the lowest triplet excitation over the (dimer) singlet ground state is apparently obtained by promoting one of the dimer singlets to a triplet. Although the dimer product remains to be the exact ground state even for non-zero $`J^{}`$, the above completely localized triplet does not; perturbation $`J^{}`$ โbroadensโ the triplet by exciting nearby singlets. Unlike the ground state which is perfectly free from quantum fluctuation, excited states strongly suffer from it. This is one of the most important features of the model. As a result, (physical) triplets can interact with each other with the help of virtual triplets created by perturbation and the effective Hamiltonian for the physical triplet degrees of freedom should contain effective interactions.
The most systematic way to take into account such virtual processes would be the strong-coupling expansion. We start from the limit $`J^{}=0`$ and treat the interaction with coupling $`J^{}`$ by perturbation. In the absence of the external field, the spin states of a single isolated dimer ($`J`$) bond consists of a singlet and triplet separated by a gap $`J`$. When the field is increased until the lowest triplet with $`S^z=1`$ intersects the singlet, we may keep only two statesโthe singlet and the lowest tripletโas the physical degrees of freedom as far as the low-energy sector is considered.
We carry out the degenerate perturbation for such low-energy degrees of freedom. Considering the triplet ($`S=1`$) state with $`S^z=1`$ as a particle (a hard-core boson) and the dimer singlet ($`S=0`$) as a vacancy, we derive an effective Hamiltonian for the magnetic particles. (Rest of spin states, i.e. triplets with $`S^z=0`$ and $`1`$, are included into the intermediate virtual states of the perturbation.) The perturbational expansion is performed up to the 3rd order in $`J^{}/J`$ for degenerate states with a constant number of dimer triplet excitations with $`S^z=1`$. The final form of the effective Hamiltonian is as follows
$`H`$ $`=`$ $`H_0+H_1+H_2+H_3,`$ (1)
$`H_0`$ $`=`$ $`(JB){\displaystyle \underset{i}{}}n_i,`$ (2)
$`H_1`$ $`=`$ $`{\displaystyle \frac{J^{}}{2}}{\displaystyle \underset{i,j}{}}n_in_j,`$ (3)
$`H_2`$ $`=`$ $`{\displaystyle \frac{J^2}{J}}{\displaystyle \underset{i}{}}n_i+{\displaystyle \frac{J^2}{2J}}{\displaystyle \underset{i,j}{}}n_in_j`$ (4)
$`+`$ $`{\displaystyle \frac{J^2}{4J}}{\displaystyle \underset{iA}{}}\{2n_{i+\text{e}1}(1n_i)n_{i\text{e}1}+(b_{i+\text{e}2}^{}b_{i\text{e}2}+h.c.)n_i`$ (6)
$`+(b_i^{}b_{i+\text{e}2}b_i^{}b_{i\text{e}2}+h.c.)(n_{i\text{e}1}n_{i+\text{e}1})\}`$
$`+`$ $`{\displaystyle \frac{J^2}{4J}}{\displaystyle \underset{iB}{}}\{\text{e}_1\text{e}_2\},`$ (7)
$`H_3`$ $`=`$ $`{\displaystyle \frac{J^3}{2J^2}}{\displaystyle \underset{i}{}}n_i{\displaystyle \frac{J^3}{8J^2}}{\displaystyle \underset{\mathrm{NN}}{}}n_in_j+{\displaystyle \frac{J^3}{4J^2}}{\displaystyle \underset{\mathrm{NNN}}{}}n_in_j`$ (8)
$`+`$ $`{\displaystyle \frac{J^3}{16J^2}}{\displaystyle \underset{iA}{}}[4n_{i+\text{e}2}n_in_{i\text{e}2}12n_{i+\text{e}1}(n_i1)n_{i\text{e}1}`$ (32)
$`+6(b_{i+\text{e}2}^{}b_{i\text{e}2}+h.c.)n_i`$
$`+5(b_i^{}b_{i+\text{e}2}b_i^{}b_{i\text{e}2}+h.c.)(n_{i\text{e}1}n_{i+\text{e}1})`$
$`(b_{i+\text{e}2}^{}b_i+h.c.)(n_{i+\text{e}2+\text{e}1}n_{i\text{e}1}n_{i+\text{e}2\text{e}1}n_{i+\text{e}1})`$
$`(b_{i\text{e}2}^{}b_i+h.c.)(n_{i\text{e}2\text{e}1}n_{i+\text{e}1}n_{i\text{e}2+\text{e}1}n_{i\text{e}1})`$
$`2(b_{i+\text{e}2+\text{e}1}^{}b_i+h.c.)n_{i+\text{e}2}n_{i\text{e}1}`$
$`2(b_{i\text{e}2\text{e}1}^{}b_i+h.c.)n_{i\text{e}2}n_{i+\text{e}1}`$
$`2(b_{i\text{e}2+\text{e}1}^{}b_i+h.c.)n_{i\text{e}2}n_{i\text{e}1}`$
$`2(b_{i+\text{e}2\text{e}1}^{}b_i+h.c.)n_{i+\text{e}2}n_{i+\text{e}1}`$
$`+(b_{i+\text{e}2}^{}b_i+h.c.)(n_{i+\text{e}2+\text{e}1}n_{i+\text{e}1}n_{i+\text{e}2\text{e}1}n_{i\text{e}1})`$
$`+(b_{i\text{e}2}^{}b_i+h.c.)(n_{i\text{e}2\text{e}1}n_{i\text{e}1}n_{i\text{e}2+\text{e}1}n_{i+\text{e}1})`$
$`+2(b_{i+\text{e}2+\text{e}1}^{}b_i+h.c.)n_{i+\text{e}2}n_{i+\text{e}1}`$
$`+2(b_{i\text{e}2\text{e}1}^{}b_i+h.c.)n_{i\text{e}2}n_{i\text{e}1}`$
$`+2(b_{i\text{e}2+\text{e}1}^{}b_i+h.c.)n_{i\text{e}2}n_{i+\text{e}1}`$
$`+2(b_{i+\text{e}2\text{e}1}^{}b_i+h.c.)n_{i+\text{e}2}n_{i\text{e}1}`$
$`+2(b_{i+\text{e}2\text{e}1}^{}b_{i\text{e}2}b_{i+\text{e}2+\text{e}1}^{}b_{i\text{e}2}+h.c.)n_{i+\text{e}2}n_i`$
$`+2(b_{i\text{e}2\text{e}1}^{}b_{i+\text{e}2}b_{i\text{e}2+\text{e}1}^{}b_{i+\text{e}2}+h.c.)n_{i\text{e}2}n_i`$
$`(b_{i+\text{e}2}^{}b_{i\text{e}2}+h.c.)n_i`$
$`\text{ }\times (n_{i+\text{e}2+\text{e}1}+n_{i+\text{e}2\text{e}1}+n_{i\text{e}2\text{e}1}+n_{i\text{e}2+\text{e}1})`$
$`\{(b_i^{}b_{i\text{e}2}+h.c.)n_{i+\text{e}2}(b_i^{}b_{i+\text{e}2}+h.c.)n_{i\text{e}2}\}`$
$`\text{ }\times (n_{i\text{e}1}n_{i+\text{e}1})`$
$`+2(b_{i+\text{e}2}^{}b_{i\text{e}2}+h.c.)n_i(n_{i\text{e}1}+n_{i+\text{e}1})`$
$`3(b_i^{}b_{i\text{e}1}+h.c.)n_{i+\text{e}1}(n_{i\text{e}2\text{e}1}n_{i+\text{e}2\text{e}1})`$
$`3(b_i^{}b_{i+\text{e}1}+h.c.)n_{i\text{e}1}(n_{i+\text{e}2+\text{e}1}n_{i\text{e}2+\text{e}1})]`$
$`+`$ $`{\displaystyle \frac{J^3}{16J^2}}{\displaystyle \underset{iB}{}}[\text{e}_1\text{e}_2],`$ (33)
where indices $`i`$ and $`j`$ run over an effective square lattice of dimer bonds (both horizontal and vertical), and $`A`$ ($`B`$) sublattice contains horizontal (vertical) ones. The operator $`b_i^{}`$ ($`b_i`$) creates (annihilates) magnetic particle with spin $`S^z=1`$ at bond $`i`$, and $`n_i=b_i^{}b_i`$. The unit vectors $`\text{e}_1`$ and $`\text{e}_2`$ are shown in Fig. 1. The interactions summed up on B sublattices are obtained by replacing $`\text{e}_1`$ and $`\text{e}_2`$ in those on A sublattices. The abbreviations NN and NNN denote pairs of nearest neighbor and next nearest neighbor sites.
The Hamiltonian derived above does not have the bare one-particle hopping terms like $`b_i^{}b_j+b_j^{}b_i`$, which means a single dimer triplet excitation is dispersionless at this order. This is in a striking contrast with other spin gap systems e.g. the 2-leg ladder. The energy gap of one dimer triplet excitation is evaluated as $`\mathrm{\Delta }E_1(B)=JBJ^2/JJ^3/(2J^2)`$.
On the other hand, the effective Hamiltonian contains many correlated-hopping processes, where an effective hopping of a particle is mediated by another one. Roughly speaking, these are closely related to 2-particle Green functions of the triplet bosons. This correlated hopping is important for dynamics of the Shastry-Sutherland model and it leads to many interesting conclusions; bound states without attraction, supersolid with stripe structure, etc. Most terms of higher orders concern the correlated hopping.
Many two-body repulsive interactions are also derived, whose range and geometry are shown in Fig. 4 of Ref. . The nearest neighbor repulsion $`V_{\mathrm{NN}}`$ and the next-nearest-neighbor one $`V_{\mathrm{NNN}}`$ are derived as $`V_{\mathrm{NN}}=J^{}/2+J^2/2JJ^3/8J^2`$ and $`V_{\mathrm{NNN}}=J^3/4J^2`$. (There were a typographical error and a mistake in the coefficients of the 3rd order terms of $`V_{\mathrm{NN}}`$ and $`V_{\mathrm{NNN}}`$ in ref. . There was also the same mistake in ref. .) The 3rd-neighbor repulsion $`V_{3\mathrm{r}\mathrm{d}}`$ between vertical (horizontal) bonds with the distance $`2\text{e}_1`$ ($`2\text{e}_2`$) is $`V_{3\mathrm{r}\mathrm{d}}=J^2/2J+3J^3/4J^2`$. Thus $`V_{3\mathrm{r}\mathrm{d}}`$ is anisotropic and acts only in one direction. The effective Hamiltonian (1) correctly reflects the space-group symmetry of the original Shastry-Sutherland model. If it is considered as a model of interacting hard-core bosons on a square lattice consisting of dimer bonds, it still contains matrix elements which are not invariant under naive $`90^{}`$ rotation about one site. We show in section III that they lead to anisotropic SDW states with a stripe structure. Longer-range repulsions between particles can appear from higher-order perturbations. Actually, Miyahara and Ueda independently took into account such long-range interactions in a phenomenological manner, also adding weak one-body hopping by hand. Our finding is that there is much stronger correlated hopping processes and that triplets are not necessarily localized.
## III Magnetization process
In order to discuss the ground state properties of the effective Hamiltonian (1), we first consider only repulsive interactions. From naive consideration of the range and geometry of the repulsive interactions (see Fig. 4 in Ref. ), we can imagine various insulating density-wave states of dimer triplets. Let us treat the repulsive interactions obtained at different orders of $`J^{}/J`$, separately.
* The strongest interaction comes from the 1st-order term and it is repulsive for adjacent triplets. This repulsive interaction chooses SDW with 2-(dimer)sublattices checkerboard structure at $`m/m_{\mathrm{sat}}=1/2`$, where the unit cell has 4 sites in the original Shastry-Sutherland lattice. (See Fig. 6(b) in Ref. .)
* The next strong repulsive interactions originate from 2nd order terms and they are contained in $`V_{\mathrm{NN}}`$ and $`V_{\mathrm{NNN}}`$. Because of anisotropic interaction $`V_{3\mathrm{r}\mathrm{d}}`$, a stripe SDW state with a 6-(dimer)sublattice structure is stabilized at $`m/m_{\mathrm{sat}}=1/3`$ as shown in Fig. 4.
* The 3rd-order perturbation generates the weakest interactions, which favor SDW of 10-(dimer)sublattices checkerboard structure at $`m/m_{\mathrm{sat}}=1/5`$.
These SDW states do not necessarily appear if we consider all interactions and correlated hopping terms together. Whether SDW insulator is realized or not is determined by competition between repulsive interactions and correlated hopping.
To consider both effects of correlated hoppings and interactions together, we study the effective Hamiltonian in the classical limit. To this end, we map the hard-core boson system to the $`S=1/2`$ quantum (pseudo)spin system and then approximate the Pauli matrices by the components of a classical unit vector. Within the mean-field approximation, we first search for the ground state taking into account two and four (dimer)sublattices with a checkerboard structure and also three (dimer)sublattices with a stripe structure, because insulating states with these configurations are expected from the above consideration. To take account of larger sublattice structures, we next study the ground state of a finite system with a Monte Carlo method, where we gradually decrease temperatures to zero. We consider a finite cluster of 60 dimers and impose a periodic boundary condition that matches with all SDW configurations expected from the repulsive interactions, i.e. 2-(dimer)sublattice and 5-(dimer)sublattice checkerboard structures and 3-(dimer)sublattice stripe one.
The evaluated magnetization processes are shown for the cases $`J/J^{}=0.45`$, $`0.6`$, and $`0.68`$ in Figs. 2(a), (b), and (c), respectively. Plateau structures appear at magnetization $`m/m_{\mathrm{sat}}=1/3`$ and 1/2, but no plateau appears at $`m/m_{\mathrm{sat}}=1/5`$ and 1/4. Near $`m/m_{\mathrm{sat}}=1/4`$, the slope of the curve becomes less steep, but not flat. This means that the $`m/m_{\mathrm{sat}}=1/4`$ state is energetically stable, though it does not have spin gap. In the plateau phases at $`m/m_{\mathrm{sat}}=1/3`$ and 1/2, there are SDW orders and no off-diagonal long-range order (ODLRO), that is, collinear spin states are realized. In the non-plateau phases, spins have ODLRO. Spin configuration of each phase is explained in the following subsections. A remark is in order here about our classical approximation. We obtained several plateaus by analyzing the classical pseudo-spin Hamiltonian. From this, one may conclude that these plateaus are of a classical origin. However, the spins approximated by vectors are not the original ones but are pseudo spins obtained for quantum objectsโspin singlet and tripletโand these plateaus are not classical.
As has been mentioned above, the spin configurations obtained in our approximate method concern the pseudo spin defined by triplet ($`S^z=+1`$) and singlet on a dimer bond. In order to translate the classical pseudo-spin configuration into the original $`S=1/2`$ one, it is convenient to consider the so-called spin-coherent state which realizes almost โclassicalโ states using quantum states. It is well-known that an $`S=1/2`$ coherent state specified by a classical unit vector
$$\stackrel{}{\mathrm{\Omega }}=(\mathrm{cos}\mathrm{\Phi }\mathrm{sin}\mathrm{\Theta },\mathrm{sin}\mathrm{\Phi }\mathrm{sin}\mathrm{\Theta },\mathrm{cos}\mathrm{\Theta })$$
(34)
is given by (aside from a phase factor coming from the gauge degrees of freedom)
$$|\mathrm{\Phi },\mathrm{\Theta }=\text{e}^{i\frac{1}{2}\mathrm{\Phi }}\mathrm{sin}\left(\frac{\mathrm{\Theta }}{2}\right)|\text{singlet}+\text{e}^{i\frac{1}{2}\mathrm{\Phi }}\mathrm{cos}\left(\frac{\mathrm{\Theta }}{2}\right)|\text{triplet}.$$
(35)
Plugging the expressions of $`|\text{singlet}`$ and $`|\text{triplet}`$ in terms of original $`S=1/2`$, we obtain
$`|\mathrm{\Phi },\mathrm{\Theta }`$ $`=`$ $`\text{e}^{i\frac{1}{2}\mathrm{\Phi }}\mathrm{sin}\left({\displaystyle \frac{\mathrm{\Theta }}{2}}\right){\displaystyle \frac{1}{\sqrt{2}}}(||)`$ (37)
$`+\text{e}^{i\frac{1}{2}\mathrm{\Phi }}\mathrm{cos}\left({\displaystyle \frac{\mathrm{\Theta }}{2}}\right)|.`$
Taking expectation values
$`\mathrm{\Phi },\mathrm{\Theta }|S_1^+|\mathrm{\Phi },\mathrm{\Theta }`$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{2}}}\text{e}^{i\mathrm{\Phi }}\mathrm{sin}\mathrm{\Theta }`$ (38)
$`\mathrm{\Phi },\mathrm{\Theta }|S_2^+|\mathrm{\Phi },\mathrm{\Theta }`$ $`=`$ $`+{\displaystyle \frac{1}{2\sqrt{2}}}\text{e}^{i\mathrm{\Phi }}\mathrm{sin}\mathrm{\Theta }`$ (39)
$`\mathrm{\Phi },\mathrm{\Theta }|S_1^z|\mathrm{\Phi },\mathrm{\Theta }`$ $`=`$ $`\mathrm{\Phi },\mathrm{\Theta }|S_2^z|\mathrm{\Phi },\mathrm{\Theta }={\displaystyle \frac{1}{4}}(1+\mathrm{cos}\mathrm{\Theta }),`$ (40)
we can see that in the presence of the superfluid LRO ($`\mathrm{\Theta }0`$, $`\pi `$) the off-diagonal elements of the original $`S=1/2`$ spins on a dimer bond align in an antiparallel manner.
In order to check the accuracy of the strong coupling expansion and the classical approximation, we also studied a finite system of the original Shastry-Sutherland model using exact diagonalization method. We diagonalized 24-sites system with a periodic boundary condition (Fig. 3) that matches with the 12-sublattice structure at magnetization above $`m/m_{\mathrm{sat}}=1/3`$ (see section III B). The results are shown in Fig. 2. Total behavior shows good agreement with the results from the strong coupling expansion.
In the following, we discuss the nature of each phase and phase diagram.
### A SDW on plateau states
On the plateau states, dimer triplet excitations crystallize forming SDW long-range orders, and there are spin gaps. The plateau states at $`m/m_{\mathrm{sat}}=1/3`$ and 1/2 have SDW with stripe and checkerboard structures, respectively, which are shown in Figs. 6(a) and (b) in Ref. . These configurations are consistent with the insulating states naively expected from the range and geometry of repulsive interactions. The particles are perfectly closed packed at $`m/m_{\mathrm{sat}}=1/2`$ and 1/3 avoiding the repulsive interactions from 1st- and 2nd-order perturbations, respectively. (See Fig. 4 for the case of $`m/m_{\mathrm{sat}}=1/3`$.)
### B Supersolid on non-plateau states
By applying a stronger magnetic field than the critical value, the plateau states continuously change to supersolid states in which superfluid components appear and coexist with SDW of the plateau states. Since the appearance of the superfluid component is accompanied by the Goldstone bosons, the plateau gap collapses.
Above the magnetization $`m/m_{\mathrm{sat}}=1/3`$, particles with density 1/3 still form SDW with the stripe structure, which equals to the one at $`m/m_{\mathrm{sat}}=1/3`$, and the rest of particles become superfluid in โcanalsโ between stripes. (See Fig. 5(a).) Particles can hop across a line of SDW with the help of correlated hopping terms (see Fig. 6) and this hopping makes correlations between superfluids in canals. These extra particles can Bose-condense and the phases of the superfluid particles align ferromagnetically inside an individual canal and antiferromagnetically between canals.
The supersolid state above $`m/m_{\mathrm{sat}}=1/2`$ maintains the same SDW as the $`m/m_{\mathrm{sat}}=1/2`$ case and have superfluid components as well. Phases of superfluid components form a stripe structure and the phase of one stripe aligns antiparallelly to those of next ones. (See Fig. 5(b).)
In both supersolid phases, a single dimer triplet can move by itself assisted by SDW LRO and it behaves as an elementary particle of the superfluid. The SDW forms a global network of crystallized particles and it helps a dimer triplet to hop along the network owing to the correlated hopping. We roughly evaluate the hopping of an extra particle by treating the stripe SDW as a classical background. In Fig. 6, we show the matrix element of single-dimer-triplet hopping along a stripe. Note that triplet excitations are not confined in a one-dimensional space (โcanalโ) between two rows of particles. A dimer triplet can hop both parallel and perpendicular to the stripe. The superfluid component of the system may behave as an anisotropic two-dimensional Bose gas. The hopping matrices inside of a stripe are negative, whereas those across a line of triplet excitations are positive. This makes the phases of superfluid component ferromagnetic inside the stripe and antiferromagnetic between stripes.
A similar anisotropic superfluid was reported for the Bose-Hubbard model with frustrating interactions .
### C Phase diagram
We show phase diagram for $`J^{}/J`$ vs. $`B/J`$ at zero temperature in Fig. 7, where the phase boundaries are not very accurate. The plateau at $`m/m_{\mathrm{sat}}=1/2`$ appears only in the region $`0<J^{}/J<0.51`$, and one at $`m/m_{\mathrm{sat}}=1/3`$ in $`0<J^{}/J<0.95`$. For large $`J^{}/J`$, insulating phases disappear and become superfluid phases. This is because correlated hoppings are dominant in the higher order terms and they become efficient for large $`J^{}/J`$. The phase diagram may be not qualitatively accurate for large $`J^{}/J`$, since our arguments are based on a strong coupling expansion. Furthermore there is a possibility that the elementary particles change to plaquette triplets for large $`J^{}/J`$, which is not taken into account in our approximation.
Spin configurations at various values of magnetization are summarized as follows:
* For $`0<m/m_{\mathrm{sat}}<1/3`$,
spin configuration has large sublattice structures. The system is highly frustrated and we sometimes reached to different ground states with different magnetizations in the Monte Calro method. We expect that weak interactions or quantum fluctuations can drastically change the ground state properties. One reason for this ambiguity may be that a single dimer triplet is not an elementary particle at very low magnetization, but a bound pair of triplets is. (See section IV)
* At $`m/m_{\mathrm{sat}}=1/3`$,
there is only SDW LRO with stripe structure. (Fig. 4)
* For $`1/3<m/m_{\mathrm{sat}}<m_{c1}`$ with $`0.5<J^{}/J<0.8`$, and for $`1/3<m/m_{\mathrm{sat}}<1/2`$ with $`J^{}/J<0.5`$,
the ground state shows supersolid with stripe structure. (Fig. 5(a))
* At $`m/m_{\mathrm{sat}}=1/2`$ for $`J^{}/J<0.5`$,
there is only SDW LRO with checkerboard structure. (Fig. 6(b) in Ref. .) The results of the exact diagonalization (Figs. 2(a) and (b)) indicate that this phase seems to be more stable than in our approximation because of quantum fluctuations. The phase boundary at $`J^{}/J=0.51`$ might move to a larger value due to quantum effects. One possible reason for this discrepancy is that our classical approximation favors wave-like states (e.g. superfluid) against the localized state.
* For $`1/2<m/m_{\mathrm{sat}}<m_{c2}`$, if $`J^{}/J<0.5`$,
spins become supersolid with stripe structure. (Fig. 5(b))
* For $`m_{c1}<m/m_{\mathrm{sat}}<1`$ with $`0.5<J^{}/J`$, and for $`m_{c2}<m/m_{\mathrm{sat}}<1`$ with $`J^{}/J<0.5`$,
large sublattice structures, e.g. a helical structure, appear. We will discuss nearly saturated region in Section V. A new quantum phase may appear in this region.
## IV Bound state of two dimer triplets near $`m=0`$
In this section, we consider a striking effect of correlated hopping on the dynamics of the Shastry-Sutherland model. It is most clearly seen in the low-magnetization region where the number of magnetically excited triplets is small.
First we suppose that there are only two excited triplets. When they are far apart from each other, an individual triplet can hardly hop and it gains little energy by moving on a lattice. (This almost localized property was actually observed in inelastic neutron-scattering experiments.) On the other hand, when the two are adjacent to each other, the situation is completely different. From the effective Hamiltonian, we can easily see that correlated-hopping processes make coherent motion of two triplets possible, where a pair of triplet dimer excitations form a bound state with $`S^z=2`$. (In the same way we can easily derive various bound states with $`S=0`$, 1, 2 at zero magnetic field. Here we only discuss the state with $`S^z=2`$, which becomes dominant under the magnetic field.)
Using the effective Hamiltonian, we can exactly show bound states. Because of the correlating hoppings, one triplet excitation exists necessarily in nearest neighbor or next nearest neighbor sites of the other. A little calculation shows that the hopping processes are decomposed into the center-of-mass motion and the relative motion, and that the latter is closed within the four states shown in Fig. 8. Hence we can write the pair excited states as
$`{\displaystyle \underset{iA}{}}\mathrm{exp}(i๐๐ซ_i)`$ $`\{`$ $`c_1(๐)|i,1+c_2(๐)|i,2+c_3(๐)|i,3`$ (42)
$`+c_4(๐)|i,4\},`$
where the two-dimensional momenta $`\text{P}=(p_x,p_y)`$ are defined with respect to the chemical unit cell of the Shastry-Sutherland lattice and unit vectors are defined in Fig. 9. (In this sense, the readers should not be confused our P with that used in refs. and .) The energy spectra $`\omega _j(\text{P})`$ ($`j=1,2,3,4`$) of them are computed by diagonalizing the following hopping matrix:
$$\left(\begin{array}{cccc}2V_0+V_{\mathrm{NNN}}& J_{\mathrm{NN}}& J_{\mathrm{NN}}\text{e}^{iP_y}& 0\\ J_{\mathrm{NN}}& 2V_0+V_{\mathrm{NN}}& J_{3rd}& J_{\mathrm{NN}}\text{e}^{iP_x}\\ J_{\mathrm{NN}}\text{e}^{iP_y}& J_{3rd}& 2V_0+V_{\mathrm{NN}}& J_{\mathrm{NN}}\\ 0& J_{\mathrm{NN}}\text{e}^{iP_x}& J_{\mathrm{NN}}& 2V_0+V_{\mathrm{NNN}}\end{array}\right),$$
(43)
where
$`V_0`$ $`=`$ $`JB{\displaystyle \frac{J^2}{J}}{\displaystyle \frac{J^3}{2J^2}},`$ (44)
$`J_{\mathrm{NN}}`$ $`=`$ $`{\displaystyle \frac{J^2}{4J}}+{\displaystyle \frac{5J^3}{16J^2}},`$ (45)
$`V_{\mathrm{NN}}`$ $`=`$ $`{\displaystyle \frac{J^{}}{2}}+{\displaystyle \frac{J^2}{2J}}{\displaystyle \frac{J^3}{8J^2}},`$ (46)
$`V_{\mathrm{NNN}}`$ $`=`$ $`{\displaystyle \frac{J^3}{4J^2}},`$ (47)
$`J_{3\mathrm{r}\mathrm{d}}`$ $`=`$ $`{\displaystyle \frac{J^2}{4J}}+{\displaystyle \frac{3J^3}{8J^2}}.`$ (48)
In addition, there is another type of correlated motion obtained from the above one by reversing the space about $`y`$ axis. Roughly speaking, this corresponds to the interchange of dimers A and B, and its spectra are given by $`\omega _j(p_x,p_y)`$. These branches together with a dispersionless band of a pair isolated triplets give the entire spectra of two-triplet sector.
Since these bound states can move because of correlated hopping process, energy spectra of these states are dispersive. We show only the lowest branch in Fig. 10. The lowest energy is given at $`\text{P}=(\pi ,\pi )`$ as
$`\mathrm{\Delta }E_2(B)`$ $`=`$ $`2V_0+{\displaystyle \frac{1}{2}}\{J_{3\mathrm{r}\mathrm{d}}+V_{\mathrm{NN}}+V_{\mathrm{NNN}}`$ (50)
$`\sqrt{\left(J_{3\mathrm{r}\mathrm{d}}V_{\mathrm{NN}}+V_{\mathrm{NNN}}\right)^2+16J_{\mathrm{NN}}^2}\}.`$
On the other hand, two independent dimer triplet excitations (scattering state) have the dispersionless energy $`2\mathrm{\Delta }E_1(B)=2V_0`$ (small dispersion appears at 6th order and higher ). Expanding the right-hand side of eq.(50) in $`J^{}/J`$, we can readily verify that there is a gain in kinetic energy of $`J(J^{}/J)^3/4`$ from the 2-particle threshold. Actually, the lowest energy of the bound state is smaller than that of two independent dimer triplets state for any $`J^{}/J`$ as shown in Fig. 11. For example, for $`J^{}/J=0.68`$ and $`B=0`$, the dispersion of a bound state takes the minimal value $`0.442J`$ at $`\text{P}=(\pi ,\pi )`$, whereas two independent dimer triplets have the energy $`0.761J`$ in total. As is easily seen, two triplets combined to form the $`S=2`$ bound states actually feel repulsive interactions between each other (i.e. $`V_{\mathrm{NN}},V_{\mathrm{NNN}}>0`$). Since strong repulsion $`V_{\mathrm{NN}}`$ acts for a pair on adjacent bonds, one may naively expect that such bound motions are not energetically favorable. However, we can take an optimal linear combination of the four relative configurations $`|i,\alpha `$ with $`\alpha =1,2,3,4`$, so that the bound state may avoid the effect of $`V_{\mathrm{NN}}`$ (note that only $`|i,2`$ and $`|i,3`$ feel the repulsion $`V_{\mathrm{NN}}`$) while gaining the kinetic energy by the coherent motion. For large $`V_{\mathrm{NN}}`$ we can easily verify that the effect of $`V_{\mathrm{NN}}`$ is canceled in $`\mathrm{\Delta }E_2`$. The energy gain due to the motion is larger than the cost from repulsion, and hence relatively stable bound states are formed.
Next, we apply a magnetic field in the $`z`$-direction. Because $`S^z=2`$ bound states have lower energy than two unbound triplets, the bottom of the bound states first touches the ground state when the field is increased and one-triplet excitation still has a finite energy gap at the critical magnetic field. By increasing the magnetic field more than the critical field, a macroscopic number of bound states condense instead of the single dimer triplets. Thus the non-plateau state at very low magnetization may be a superfluid of bound states and is different from the almost localized dimer triplet state discussed by Miyahara and Ueda. The critical magnetic field of magnetization process, where the magnetization starts to increase from zero, corresponds to half of the energy gap $`\mathrm{\Delta }E_2(0)`$. If we regard the slow increase of magnetization below $`H_1=22.5`$\[T\] as the consequence of bound state condensation, we can estimate the gap $`\mathrm{\Delta }E_2(0)`$ as $`51.0`$\[K\] from magnetization process. On the other hand the one-triplet energy gap $`\mathrm{\Delta }E_1(0)`$ is estimated as $`34.7`$\[K\] from susceptibility, inelastic neutron scattering, and ESR. If we set the parameters as $`J=81.4`$\[K\] and $`J^{}=53.5`$\[K\], estimates of the energy gaps, $`\mathrm{\Delta }E_1(0)`$ and $`\mathrm{\Delta }E_2(0)`$, coincide with the experimental results.
We expect that these bound states are destroyed at higher magnetization. Indeed, one-triplet excitations can move around above $`m/m_{\mathrm{sat}}=1/3`$, as discussed in section III B, and they can gain more kinetic energy than bound states. There must be a transition of elementary particles from bound states for very low magnetization to one-triplet states for high magnetization. We can see this transition, if we neglect interactions between particles and apply mean-field approximation to the correlated hopping term. As particle density is increased, one-particle hopping process effectively appears as $`(b_i^{}b_j+\text{H.c.})n`$ from the correlated hopping term and then triplet particles gain kinetic energy. In this rough estimation, a pair of unbound triplets have lower energy than the bound state above $`m/m_{\mathrm{sat}}=0.253`$ for $`J^{}/J=0.63`$. We can expect that bound states disintegrate above a finite value of magnetization and one-triplet particles turn to elementary particles. This estimate of critical magnetization can be highly modified by strong correlation and readers should not consider the above value seriously.
Finally we mention about crystallization of bound states. One may expect bound states to crystallize at low magnetization, but it will not occur. If bound states tend to crystallize, bound states loose kinetic energy and then they will be unbound by repulsive interaction between triplets.
## V Near Saturation
There is another region where quantum effects manifest themselves. In this section, we briefly discuss the region just below the saturation field. In this region the one-particle excitation can be obtained from the original Shastry-Sutherland model without any approximation.
The single-particle excitation over the fully polarized ground state is given by a single flipped spin. The dispersion of this excitation is readily computed by diagonalizing the following hopping matrix
$$\left(\begin{array}{cccc}\frac{1}{2}(J+4J^{})& \frac{J^{}}{2}(1+\text{e}^{ip_x})& \frac{J}{2}\text{e}^{ip_x}\text{e}^{ip_y}& \frac{J^{}}{2}(1+\text{e}^{ip_y})\\ \frac{J^{}}{2}(1+\text{e}^{ip_x})& \frac{1}{2}(J+4J^{})& \frac{J^{}}{2}(1+\text{e}^{ip_y})& \frac{J}{2}\\ \frac{J}{2}\text{e}^{ip_x}\text{e}^{ip_y}& \frac{J^{}}{2}(1+\text{e}^{ip_y})& \frac{1}{2}(J+4J^{})& \frac{J^{}}{2}(1+\text{e}^{ip_x})\\ \frac{J^{}}{2}(1+\text{e}^{ip_y})& \frac{J}{2}& \frac{J^{}}{2}(1+\text{e}^{ip_x})& \frac{1}{2}(J+4J^{})\end{array}\right).$$
(51)
In the above equation, momenta $`(p_x,p_y)`$ are defined with respect to the chemical unit cell. Reflecting the fact that a single unit cell contains four spins, the spectrum consists of four bands. Note that the four eigenvalues are invariant under the point group $`\text{D}_{2\mathrm{d}}`$:
$$(p_x,p_y)(p_y,p_x)\text{and}(p_x,p_y)(p_x,p_y).$$
(52)
In the dimer limit $`J^{}/J1`$, the lower two correspond to energy of a singlet particle on dimer bonds, where the number two comes from the two mutually orthogonal dimer bonds, and the higher two to a triplet ($`S^z=0`$) one; our approximation in section II corresponds to neglecting the latter as higher-lying.
Although the expression of the dispersion is rather complicated, we can locate the position of the band minima in the momentum space, which is relevant in determining the structure in the vicinity of saturation. For $`0J^{}/J<1`$, the lowest band takes a minimal value $`J(1+J^{}/J)^2`$ on a closed curve
$$\mathrm{cos}p_x+\mathrm{cos}p_y=2\left(\frac{J^{}}{J}\right)^2.$$
(53)
This implies that magnetization saturates at
$$H_\mathrm{c}=J\left(1+\frac{J^{}}{J}\right)^2.$$
(54)
Because of a dispersionless mode on the closed curve, the density of states (DOS) near saturation magnetization is like 1D and different from the usual 2D one. Hence the singularity of magnetization curve near saturation magnetization would be 1D-like, i.e., $`|m_\mathrm{c}m|\sqrt{|H_\mathrm{c}H|}`$. This behavior can be seen in the results of the exact diagonalization in Fig. 2.
The location of the dispersion minimum $`๐_{\mathrm{min}}`$ together with the corresponding eigenvector determines the spin structure at the semiclassical level. Usually spin states just below the critical field are correctly given by the classical model . Detailed analyses of the wave functions reveal that the fourfold-degenerate classical helical order, which was pointed out by Shastry and Sutherland, corresponds to the four apexes $`(\pm p_{\mathrm{max}},0),(0,\pm p_{\mathrm{max}})`$ ($`p_{\mathrm{max}}=2\mathrm{cos}^1(J^{}/J)`$) of the closed curve. The fate of the classical helical LRO when the quantum effects are fully taken into account is unclear at present. The four ground states with helical LRO are connected by a gapless line, which means stiffness of helical order is vanishing. We hence expect that quantum fluctuations destroy the classical LRO.
On the other hand, for $`J^{}/J1`$ the saturation field $`H_\mathrm{c}=4J^{}`$ becomes $`J`$-independent and the band minimum shrinks to a single point $`๐=(0,0)`$; the spin structure realized in the vicinity of saturation is the classical staggered one where spins connected by $`J^{}`$-bonds align antiferromagnetically in the $`xy`$-plane. Correspondingly, the transition to saturation is the same as that of 2D superfluid.
## VI Critical Phenomena
According to an analogy to many-particle theories, a plateau state corresponds to a SDW insulating state and gapless ones to supersolids. As the plateau states at $`m/m_{\mathrm{sat}}=0`$, $`1/3`$ and $`1/2`$ collapse by increasing the magnetic field, superfluid components appear, whereas the SDW structure still remains in each phase. Contrary to one dimension, superfluid LRO is accompanied by the gapless Goldstone mode in our two-dimensional case. Hence, collapse of plateau occurs at the same point with the onset of superfluidity component. On the other hand if we look at the lower phase boundaries of the SDW phases, SDW structures change at the transition points in our approximation. In this case the phase transitions are presumably of first order.
Now let us consider the case of 2nd order transition. We can imagine two different situations. The first one is (i) the transitions driven by changing the external field while the coupling $`J^{}/J`$ is kept fixed. This transition is seen in the actual magnetization process. This type of field-induced transitions has a close relationship to the filling-control insulator-to-superfluid transitions in Bose systems, and the methods used there can be imported to our case. The second one is (ii) continuous quantum phase transitions occurring with fixed values of magnetization. As the limit $`J^{}/J\mathrm{}`$ is approached, the Shastry-Sutherland model reduces to the ordinary $`S=1/2`$ square-lattice Heisenberg antiferromagnet, where no magnetization plateau appears. Hence magnetization plateaus vanish at some critical values of $`J^{}/J`$ and are superseded by supersolid (or, superfluid) phases. In the present model, there is no particle-hole symmetry around insulating phases apparently. We hence conclude that the above two transitions have the same universality class.
First of all, we have to keep in mind that because of the special geometry of the Shastry-Sutherland model there is no a priori reason for believing that the system is described by the ordinary Bose liquid with a well-defined one-particle dispersion $`\epsilon (p)p^2`$. Actually, the low-order effective Hamiltonian (1) lacks the one-particle part. However, we have seen in section IV that this leads to the formation of dispersive two-triplet bound states in the low-field region, and in section III B that SDW structure makes one particle dispersive in the supersolid phase around $`m/m_{\mathrm{sat}}=1/3`$ and 1/2.
Near the phase boundaries, the superfluid amplitude is small and we can map the problem onto the effective Ginzburg-Landau model described by the superfluid order parameter . (These Bose particles are not necessarily dimer triplet excitations, but they can be plaquette triplet states of two dimer triplets or flipped spin states.) This enables us to conclude that superfluid-onset transition would be described by the (2+1)-dimensional classical XY model when the particle-hole symmetry exists and by the $`z=2`$ mean-field-like transitions when it does not. In our case, the effective boson model in section II does not have particle-hole symmetry and we hence conclude that the plateau transition is of the dynamical exponent $`z=2`$ and behaves as
$`|mm_\mathrm{c}|`$ $``$ $`|HH_\mathrm{c}|\mathrm{log}(C/|HH_\mathrm{c}|)`$ (56)
(2D system).
in two dimensions. Here $`m_\mathrm{c}`$ and $`H_\mathrm{c}`$ denote the critical magnetization and field at the plateau transition. In the real material, there are weak interactions between two-dimensional layers and these interactions will push the system above the upper critical dimension, i.e.
$`|mm_\mathrm{c}|`$ $``$ $`|HH_\mathrm{c}|`$ (58)
(3D and quasi-2D systems).
Note that these forms are quite different from that in one-dimensional systems, $`|mm_\mathrm{c}|\sqrt{|HH_\mathrm{c}|}`$.
Here we also give a comment about the case the kinetic term of Bose particles does not behave as $`\epsilon (p)p^2`$. According to the scaling argument of Ref. , exponents of this kind of transitions should satisfy the relation
$$z\nu =1.$$
(59)
If the one-particle dispersion is of the form $`\epsilon (p)p^l`$ ($`l`$ denotes an even integer), the standard power-counting argument shows that the upper critical dimensions are given by $`d_\mathrm{u}=l`$. For $`dl`$, we obtain $`|mm_\mathrm{c}||HH_\mathrm{c}|^{d/l}`$.
For the case that SDW structure changes at phase boundary, e.g. the boundary between the plateau state at $`m/m_{\mathrm{sat}}=1/2`$ and the supersolid one below $`m/m_{\mathrm{sat}}=1/2`$, the system shows a 1st order phase transition in our approximation. We donโt find any incommensurate phase between them. To take into account the possibility of any 2nd-order phase transition or incommensurate phase, we need to consider the effect of quantum fluctuations more seriously.
## VII Discussions and future problems
To summarize, we studied the magnetic behavior of the Shastry-Sutherland model using strong-coupling expansion. Magnetic excitations show insulator-supersolid transitions at magnetization $`m/m_{\mathrm{sat}}=1/3`$ and 1/2, and thereby create magnetization plateaus. Magnetization curve obtained near $`m/m_{\mathrm{sat}}=1/3`$ looks similar to the experimental result.
At zero magnetization, bound states of triplet excitations are formed by the correlated hopping process. Above the critical field, quintuplet ($`S=2`$) bound states become elementary particles in the ground state and they condense. Whereas, for large magnetization, the bound states are destroyed in the ground states and triplet excitations become elementary particles. Triplet excitations are essential for the plateau transitions at $`m/m_{\mathrm{sat}}=1/3`$ and 1/2. In the experiments, it is unclear in which region bound states appear as elementary particles. One possibility is that they appear in the tail of the magnetization process below $`m/m_{\mathrm{sat}}0.025`$. In this region, slope of the magnetization curve is different from that of rest parts. Further detail analysis are needed on the nature of quasiparticles.
In the present analysis, we did not find the plateaus at $`m/m_{\mathrm{sat}}=1/8`$ and 1/4. The mechanism of stabilizing these plateaus is not yet clear. It may be natural to believe that dimer triplet excitations are crystallized by longer-range repulsive interaction. ($`S=2`$ bound states cannot crystallize because of a special origin of the binding energy as we discussed in section IV.) We can consider two origins for the repulsions as follows:
* Though we cut the perturbation series at 3rd order, the higher-order expansions can produce longer-range repulsions between particles. These repulsive interactions may induce crystallization at low magnetization.
* Longer-range repulsions can come from other antiferromagnetic spin interactions in the original spin model, which have not been accounted for in the pure Shastry-Sutherland model. If we treat the antiferromagnetic interactions between next-neighbor dimer bonds, we can produce crystallization at low magnetization. For example, Mรผller-Hartmann et al. demonstrated the appearance of 1/4-plateau considering another spin interaction, which acts between nearest-neighbor dimers.
It is unclear which spin interaction is important in the real material. We need first-principle calculation to estimate exchange couplings. We also need to keep in mind that the real material is close to the plaquette singlet phase. Under the magnetic field, if this phase becomes more stable than the dimer singlet state and plaquette triplets become elementary particles, plaquette triplets may crystallize and hence create magnetization plateaus at low magnetization as discussed in refs. . This should be considered in a future.
###### Acknowledgements.
We would like to thank the late Dr. Nobuyuki Katoh for stimulating discussions at the beginning of this study. We also thank Hiroshi Kageyama, Norio Kawakami, Kenn Kubo, Hiroyuki Nojiri, and Kazuo Ueda for useful comments and discussions. We also acknowledge Hiroshi Kageyama and Hiroyuki Nojiri for showing us their experimental results before publication. One of us (TM) acknowledges the condensed matter theory group in the Harvard University for kind hospitality. KT is financially supported by Inoue fellowship.
## A Effective Hamiltonian
In this appendix, we briefly explain how to obtain the effective Hamiltonian in the framework of degenerate perturbation. We suppose that the ground states of the unperturbed Hamiltonian ($`_0`$) are degenerated and we diagonalize these degenerate ground states by degenerate perturbation. Let the operators $`P_0`$ and $`Q_0`$ be the projection operators onto the degenerate (unperturbed) ground-state sector and its complement, respectively. In addition, we define a projection $`P`$ onto the perturbed ground-state sector.
According to ref. , the problem of degenerate perturbation reduces to solving the following problem
$$P_0PP_0|\text{G.S.; }\alpha =E_\alpha K|\text{G.S.; }\alpha ,$$
(A1)
where the hermitian operator $`K`$ is defined by
$$KP_0PP_0.$$
(A2)
However, this form is not so convenient to our purpose because it does not take the form of the ordinary eigenvalue problem. A trick invented by Bloch solves this difficulty. The key is to introduce a new operator $`U`$ through the following relation
$$PP_0=UK.$$
(A3)
The operator $`U`$ can be expanded as
$$U=\underset{n=0}{\overset{\mathrm{}}{}}\lambda ^nU^{(n)},$$
(A4)
where the $`n`$-th order coefficients $`U^{(n)}`$ are given by
$`U^{(0)}`$ $`=`$ $`P_0`$ (A5)
$`U^{(1)}`$ $`=`$ $`SVP_0`$ (A6)
$`U^{(2)}`$ $`=`$ $`\left(SVSV+S^2VS^0V\right)P_0`$ (A7)
$`U^{(3)}`$ $`=`$ $`(SVSVSV+S^3VS^0VS^0V+S^2VSVS^0V`$ (A10)
$`+SVS^2VS^0V+S^2VS^0VSV)P_0`$
$`\mathrm{}`$
In the above equation, we have used a short-hand notation $`S^k`$ defined by
$$S^k\{\begin{array}{cc}P_0\hfill & \text{ for }k=0\hfill \\ \frac{1}{(E_0)^k}Q_0\hfill & \text{ for }k1.\hfill \end{array}$$
(A11)
Then, eq. (A1) can be recasted as
$$\left(P_0UE_\alpha \right)K|\text{G.S.; }\alpha =0.$$
(A12)
This is a usual eigenvalue problem and the matrix we have to diagonalize is finally given by
$`_{\mathrm{eff}}`$ $`=`$ $`P_0U`$ (A13)
$`=`$ $`E_0+\lambda P_0VP_0+\lambda ^2P_0V{\displaystyle \frac{1}{E_0_0}}Q_0VP_0`$ (A16)
$`+\lambda ^3[P_0V{\displaystyle \frac{1}{E_0_0}}Q_0V{\displaystyle \frac{1}{E_0_0}}Q_0VP_0`$
$`P_0V{\displaystyle \frac{1}{(E_0_0)^2}}Q_0VP_0VP_0]+\mathrm{}.`$
Note that non-hermitian terms appear in general when we proceed to terms higher than second order. Reality of the eigenvalues is no longer guaranteed. To remedy this shortcomings, we have used the average of $`_{\mathrm{eff}}`$ and $`_{\mathrm{eff}}^{}`$, which is now hermitian.
|
warning/0006/astro-ph0006275.html
|
ar5iv
|
text
|
# Constraints from Dynamical Friction on the Dark Matter Content of Barred Galaxies
## 1 Introduction
The flatness of disk galaxy rotation curves, particularly outside the optical radius, is generally interpreted as evidence that they are embedded in massive dark matter (DM) halos. Since the mass-to-light ratio of the visible material is uncertain, the central DM density is also uncertain; even the best determined 1-D rotation curve is consistent with almost any disk mass up to a maximum that does not over-fit the inner part (van Albada et al. 1985). This degeneracy introduces a serious uncertainty into studies of galaxy formation and evolution. Here we present an argument that DM halos have the lowest possible central density consistent with not being hollow. Following van Albada & Sancisi (1986), we refer to such minimum halo models as โmaximum disks.โ
Strong bars are seen in optical images of roughly 30% of all disk galaxies (Sellwood & Wilkinson 1993) and this fraction rises to over 50% in the near IR (Eskridge et al. 2000). Their presence makes them ideal probes of the dynamics of the central regions. The rate of rotation of a bar can be characterized by the ratio $`=D_L/a_B`$, where $`D_L`$ is the corotation radius and $`a_B`$ the bar semi-major axis. More precisely, $`D_L`$ is the distance from the center to the Lagrange point on the bar major axis where the gravitational attraction balances the centripetal acceleration in the frame rotating with the bar. Theoretical arguments (Contopoulos 1980) require $`>1`$, and there is a prejudice that $`\mathrm{}>1`$ (see Sellwood & Wilkinson 1993 for a review). We describe bars for which $`11.4`$, i.e. those for which corotation is not far beyond the barโs end, as fast.
The number of barred galaxies with measured $``$ is still quite small (Elmegreen 1996), since it requires knowledge of the barโs pattern speed, $`\mathrm{\Omega }_\mathrm{p}`$, which is hard to determine. Tremaine & Weinberg (1984a) show that $`\mathrm{\Omega }_\mathrm{p}`$ can be measured directly from observations of a tracer component that satisfies the continuity equation. To date, their method has been applied to just two galaxies: Merrifield & Kuijken (1995, see also Kent 1987) find $`=1.4\pm 0.3`$ for NGC 936 while Gerssen et al. (1999) find $`=1.15_{0.23}^{+0.38}`$ for NGC 4596. A third case should be completed soon (Debattista & Williams 2000). A less direct, but probably reliable, determination of $``$ has been made for those few galaxies in which high spatial resolution, 2-D gas kinematics have been modeled. Three cases are: $`=1.3`$ in NGC 1365 (Lindblad et al. 1996), $`=1.3`$ for NGC 1300 (Lindblad & Kristen 1996), and $`=1.2`$ for NGC 4123 (Weiner et al. 2000). A more general argument can be made on the basis of the shapes of dust lanes that $`=1.2\pm 0.2`$ (van Albada & Sanders 1982; Athanassoula 1992). Other methods to determine $``$ rely on identifying resonances, such as rings in the disk (e.g. Buta 1986; Buta & Combes 1996), but are less reliable. Thus the meager reliable measurements are mostly for galaxies of earlier type, but all indicate bars are fast; this conclusion becomes much stronger, and can be extended to later Hubble types, if the dust lane argument holds.
Weinberg (1985) predicted that dynamical friction should brake the rotation rate of a bar on a time scale short compared with the ages of galaxies if a substantial density of dark matter is present in the region of the bar. In a previous paper (Debattista & Sellwood 1998, hereafter Paper I), we confirmed this prediction for non-rotating halos with isotropic velocity distributions and concluded that DM halos must have low central densities if bars are to remain as fast as those observed. Here we extend this result to halos with anisotropic velocity dispersion tensors both with and without net rotation and show that it is not significantly altered for any reasonable velocity distribution of the dark matter (cf. Tremaine & Ostriker 1999).
## 2 Methods
We present fully self-consistent, 3-D simulations of barred disks embedded in live halos. Our simulations start from axisymmetric disk and halo models which are designed to be unstable to the formation of bars.
We adopt the Kuzโmin-Toomre disk (Kuzโmin 1956; Toomre 1963; Binney & Tremaine 1987, ยง2.2.1) because its density drops off less steeply than an exponential of the same scale length. We give the disk a Gaussian density profile in the vertical direction with a scale height $`z_\mathrm{d}`$ and sharply truncate it at some radius $`R_\mathrm{t}`$. Thus the disk density in cylindrical coordinates is
$$\rho (R,z)=\{\begin{array}{cc}\frac{f_\mathrm{d}M_{\mathrm{unit}}}{(2\pi )^{3/2}z_\mathrm{d}R_\mathrm{d}^2}\frac{\mathrm{exp}[\frac{1}{2}(z/z_\mathrm{d})^2]}{[1+(R/R_\mathrm{d})^2]^{3/2}}\hfill & RR_\mathrm{t}\hfill \\ & \\ 0\hfill & R>R_\mathrm{t}\hfill \end{array},$$
(1)
where $`R_\mathrm{d}`$ is the length scale of the disk and $`f_\mathrm{d}`$ is the fraction of mass in the disk.
We set up the initial halo from a distribution function (DF) that is computed to be in equilibrium in the presence of the adopted disk, as described in Appendix A. We adopt a lowered polytrope form for the DF:
$$f(\mathrm{x}\mathrm{x}\mathrm{x},\mathrm{v}\mathrm{v}\mathrm{v})=C(E)=C\left\{\left[2E(\mathrm{x}\mathrm{x}\mathrm{x},\mathrm{v}\mathrm{v}\mathrm{v})\right]^{n\frac{3}{2}}\left[2E_{\mathrm{max}}\right]^{n\frac{3}{2}}\right\}$$
(2)
where $`C`$ is a normalization constant and $`n`$ and $`E_{\mathrm{max}}`$ are free parameters. We emphasize that the halos generated this way are not polytropes; in particular, the density reaches zero at a finite radius for all $`n>\frac{3}{2}`$, whereas true polytropes of index $`5`$ have non-zero density everywhere. By setting $`E_{\mathrm{max}}`$ to the potential energy in the plane of the disk at some radius within the grid, we guarantee that no particles are initially on orbits that would take them off the grid. When $``$ is a function if $`E`$ only, the DF is isotropic and the halo almost spherical, but in later sections we make $``$ a function of a combination of $`E`$ and $`J_z`$ which leads to anisotropic DFs and spheroidal density distributions. The procedure for selection of particles from a DF is described in Appendix B. We set some of our halos into rotation by flipping the sign of $`L_z`$ for some fraction of the halo particles, as described in Appendix C. In some cases, we flip particles in the inner part of the halo only, with a rule (equation C2) that depends on energy in such a way that rotation declines continuously to zero near some (spherical) radius $`r_{\mathrm{rot}}`$.
We give the disk particles some random velocities with a radial dispersion, $`\sigma _u`$, set to yield a constant Toomre $`Q`$, neglecting any corrections arising from disk thickness and force softening. We then use the epicyclic approximation to set the azimuthal velocity dispersion, $`\sigma _v`$ and the equation for asymmetric drift to set the mean orbital speed. (This equation sometimes has no solution at small radii, particularly for large $`Q`$, in which case, we reduce $`\sigma _u`$.) Finally, we set the vertical velocity dispersion, $`\sigma _w`$, from the 1-D vertical Jeans equation. This disk set-up procedure is approximate, particularly for larger $`Q`$, but in practice the disk quickly adjusts to an equilibrium.
We have generally chosen $`Q=0.05`$ initially in order to hasten the formation of the bar, since we are here most interested in the evolution after this event. Models run with higher initial $`Q`$ are not qualitatively different; even though the bars are initially weaker and buckle out of the plane more, they end up somewhat longer and are braked to an even greater extent. The evolution is not significantly affected by changes to the truncation radius or by doubling or halving the initial thickness of the disk.
We imposed an initial three-fold symmetry on the models by replicating particles in sets of six in such a way as to ensure that the total momentum and components of the total angular momenta about the $`x`$\- and $`y`$-axes (the $`z`$-axis being the symmetry axis) were all zero. This prevented the model from rotating or drifting relative to our grid, which could lead to excessive and asymmetric loss of particles from the grid in our very long runs.
We adopt units in which $`G=M_{\mathrm{unit}}=R_\mathrm{d}=1`$, where $`G`$ is Newtonโs constant. The total of the disk ($`M_\mathrm{d}=f_\mathrm{d}M_{\mathrm{tot}}`$) and halo ($`M_\mathrm{h}`$) masses is $`M_{\mathrm{tot}}=[1(1+R_\mathrm{t}^2/R_\mathrm{d}^2)^{1/2}]M_{\mathrm{unit}}`$. Our unit of time is therefore $`(R_\mathrm{d}^3/GM_{\mathrm{unit}})^{1/2}`$, and our adopted time step is 0.05 in this unit.
We employ the 3-D Cartesian particle-mesh code described in Sellwood & Merritt (1994), which uses an efficient FFT-based Poisson solver (James 1977). This choice of code does limit us to computing the evolution within a fixed volume, and since we wish to retain reasonable spatial resolution within the disk, we have generally been forced to bound our halos at a small radius. We find this a reasonable price to pay, since e.g. treecode simulations of models with the same number of particles would run $`100`$ times more slowly (Sellwood 1997) which would preclude the extensive exploration of parameter space we report here. Even the two simulations reported in ยง5, which use larger grids to permit more extensive halos, run only $`12`$ times more slowly than our standard grid, and are therefore still decisively less expensive than a treecode. These performance comparisons are all based on a single processor general purpose workstation; the advantage of a grid code would be even greater on parallel computers where the field evaluation is more easily optimized. Athanassoula et al. (1998) give performance comparisons between the grid code and a machine with special-purpose hardware (GRAPE3), which clearly must depend upon the workstation adopted for comparison, but from their Figure 3, it can be seen that the grid code is competitive with their GRAPE3 machine for the grid size and particle number we employ.
The numerical parameters in the simulations we report here are summarized in Table 1. We have checked that our main conclusions are insensitive to changes in particle number, to an increase in spatial resolution or the method to determine the gravitational field, etc. In particular, the evolution of the pattern speed and bar length on finer grids, or with different geometry tracked that on our standard grid pretty well, see Debattista (1998) for details. We repeated three simulations with identical physical properties but different random seeds in the generation of particles and report the results in Figure 7. Some particles were lost from the grid (typically no more than 3% in the longest runs) almost all of which were halo particles. Naturally, these weakly bound particles carried away more than their fair share of angular momentum, which decreased by as much as 5% in the worst case. A test on a larger grid showed that the evolution is imperceptibly affected by this loss.
### 2.1 Pattern speed and Lagrange point
In order to determine the bar pattern speed, $`\mathrm{\Omega }_\mathrm{p}`$, and other properties of our models, we need a well-defined center about which to perform an expansion. Despite our careful set up, the center of our $`N`$-body system wanders from the center of our grid. Following McGlynn (1984), we define the function:
$$\omega _k(x_0,y_0,z_0)=\underset{j=1}{\overset{N}{}}\left[(x_jx_0)^2+(y_jy_0)^2+(z_jz_0)^2\right]^k$$
(3)
where $`(x_j,y_j,z_j)`$ are the coordinates of the $`j`$th particle and $`(x_0,y_0,z_0)`$ those of an expansion center. Minimizing $`\omega _k`$ with respect to $`x_0`$, $`y_0`$ and $`z_0`$ gives a centroid for the system. Setting $`k=1`$ yields the center of mass, distant particles are weighted more when $`k>1`$ while $`k<1`$ gives a centroid more sensitive to small scales. We adopt $`k=\frac{1}{2}`$, which removes the dipole moment. We determine the centroid for the total system of particles, disk and halo; the disk and halo centroids typically differed significantly only during the brief interval when the bar buckled, when the position angle of the bar was anyway not well defined. We obtain an improved estimate of the centroid position from a single Newton iteration of its old value every 20 steps, immediately before each analysis step. We have verified, at a few selected times, that further refinement results in a change in the position of the centroid by $`\mathrm{}<0.02R_\mathrm{d}`$.
We expand the instantaneous distribution of disk particles in logarithmic spirals as
$$A(m,\gamma ,t)=\frac{1}{N_\mathrm{d}}\underset{j=1}{\overset{N_\mathrm{d}}{}}\mathrm{exp}im\left(\phi _j+\mathrm{tan}\gamma \mathrm{ln}R_j\right),$$
(4)
where $`\gamma `$ is the angle between the radius vector and the ridge line of the spiral. Here, $`(R_j,\phi _j)`$ are cylindrical polar coordinates (with respect to the centroid) of the $`j`$-th particle at time $`t`$. The $`m=2`$, $`\gamma =0`$ term of this expansion gives the phase and amplitude of the bar
$$A(2,0,t)=๐\mathrm{e}^{2\mathrm{i}\phi },$$
(5)
where $`๐`$ is the bar amplitude. We calculate the monotonic angular displacement of the bar, $`\varphi (t)`$, from $`\phi `$. We estimate its derivative, $`\mathrm{\Omega }_\mathrm{p}(t)\dot{\varphi }(t)`$, by fitting a straight line to 25 consecutive data points centered at $`t`$, which smoothes out rapid fluctuations in $`\mathrm{\Omega }_\mathrm{p}`$ and yields an error estimate from the standard error in this linear fit.
Having determined $`\mathrm{\Omega }_\mathrm{p}`$, we are in a position to calculate $`D_L`$. We determine the effective force along the bar major axis in the disk plane (we average the gravitational force from both sides of the center, and use the instantaneous value of $`\mathrm{\Omega }_\mathrm{p}`$); $`D_L`$ is the distance from the center at which the net force passes through zero. We use the uncertainty in $`\mathrm{\Omega }_\mathrm{p}`$ to determine that in $`D_L`$ directly. This procedure is superior to determining the radius at which $`R\mathrm{\Omega }_\mathrm{p}`$ intersects an axisymmetric rotation curve; we have found that $`D_L`$ is larger by more than $`5`$% for strong bars when $`1`$.
It should be noted that the value of $`D_L`$ is affected by the rotation curve shape. As the bar slows, the distance from the center to the Lagrange point increases more slowly when the rotation curve declines than it would do if the rotation curve were flat. Since, in the large majority of our simulations the rotation curve does in fact drop continuously from the maximum in the disk, our reported values of both $`D_L`$ and $``$ are underestimates of the values they would have in more realistic models with flat rotation curves (see ยง5).
### 2.2 Bar semi-major axis
A bar is a straight bisymmetric distortion in the density of a disk; even Fourier components of the surface density are therefore ideal for tracing the extent of the bar. Thus, for example, Lindblad et al. (1996) and Lindblad & Kristen (1996) used the phase variations of the even Fourier components to determine the lengths of the bars in NGC 1365 and NGC 1300 respectively. Here we adopt a similar approach for determining the semi-major axis, $`a_B`$, of the bars in our simulations, using only the $`m=2`$ Fourier component. This is not always an easy measurement, particularly when the disk possesses $`m=2`$ disturbances other than the bar, such as spirals, rings surrounding the bar, and outer oval distortions.
We divide the disk into radial bins of fixed width $`0.16R_\mathrm{d}`$ and determine the amplitude and phase of the second sectoral ($`m=2`$) harmonic from the particle positions within each annular bin. We estimate the errors $`\sigma _{\mathrm{amp}}`$ and $`\sigma _{\mathrm{phs}}`$ using Monte-Carlo measurements of the phase and amplitude from synthetic samples of particles drawn from a distribution with a given $`m=2`$ amplitude. Fitting these data with a 2-D spline then yielded interpolation formulae, giving the uncertainties for the number of particles in each annulus and the measured $`m=2`$ amplitude.
Spirals are the easiest to distinguish, since they generally have a different pattern speed from the bar. Thus at a fixed radius, the peak of a spiralโs azimuthal position is usually different from that of the bar. When the inner part of the spiral lines up with the bar, however, there is no way to distinguish between it and the bar; measurements of $`a_B`$ therefore fluctuate at the beat frequency of the bar and spiral pattern speeds (this same beat frequency can be seen in measurements of $`\mathrm{\Omega }_\mathrm{p}`$). This problem becomes less severe as the evolution proceeds, because a rising $`Q`$ causes the spirals to weaken.
Oval outer disks are often perpendicular to the bar (in this sense, this outer region can be considered as an outer ring of the type R<sub>1</sub> in Butaโs classification). The number of particles in these ovals is often low, and the error bars associated with their position angle correspondingly large enough to confuse the measurement of $`a_B`$, especially late in the runs when the surface density just beyond the barโs end has been depleted. To counteract this tendency, we did not include radial bins in which the error in position angle is greater than $`80^{}`$. An example of a mildly oval outer disk can be seen in Figure 4.
At later stages in the diskโs evolution, the spirals often settle to form a ring around the bar, as can be seen in Figure 4. Their position (just outside the bar) and orientation (usually aligned with the bar) suggest they are inner rings (Athanassoula & Bosma 1985; Buta 1986; Buta & Combes 1996).<sup>1</sup><sup>1</sup>1The presence of these rings in our simulations has important ramifications for the theory of ring formation and interpretation, since rings are often considered to be gas phenomena, but our collisionless simulations have no gas. Since these elliptical rings mostly line up with the bar, they constitute an additional $`m=2`$ component locked at the barโs pattern speed which further complicates identification of the bar end. When the disk is sub-divided into annuli and the $`m=2`$ amplitude plotted as a function of radius, the ring appears as an upwards bump in the smooth decrease of amplitude from the bar. We adopt, as one estimate denoted $`a_{B}^{}{}_{1}{}^{}`$, the last radial point at which this radially binned $`m=2`$ amplitude did not deviate from a linear decrease by more than its standard error, $`\sigma _{\mathrm{amp}}`$.
We obtain a second estimate, $`a_{B}^{}{}_{2}{}^{}`$, from the phase of the $`m=2`$ moment of the disk binned as for the $`a_{B}^{}{}_{1}{}^{}`$ measurement. We estimate $`a_{B}^{}{}_{2}{}^{}`$ as the radius of the outermost bin for which the phase is constant to within the standard error in that radial bin, $`\sigma _{\mathrm{phs}}`$.
We define $`a_B`$ to be the simple average of $`a_{B}^{}{}_{1}{}^{}`$ and $`a_{B}^{}{}_{2}{}^{}`$. In practice, $`a_{B}^{}{}_{1}{}^{}`$ tended to underestimate our subjective visual impression of the bar semi-major axis (particularly at early times), while $`a_{B}^{}{}_{2}{}^{}`$ tended to overestimate it. We found that the average of these two quantities did a very good job of estimating $`a_B`$. We generously define the uncertainty in $`a_B`$, which is not a formal error, as half the difference between $`a_{B}^{}{}_{1}{}^{}`$ and $`a_{B}^{}{}_{2}{}^{}`$. Because of the formation of rings and ovals, this uncertainty tends to be largest at late times, and can be as large as several disk scale-lengths.
The uncertainty in $`D_L`$ is always much less than our generous estimates of the uncertainty in $`a_B`$, which therefore dominates our quoted uncertainty in $``$.
### 2.3 Parameterizing rotation curve decompositions
One estimator of the relative contributions of the disk and halo to the central attraction is the parameter
$$\eta \left(\frac{V_{\mathrm{c},\mathrm{disk}}}{V_{\mathrm{c},\mathrm{halo}}}\right)^2|_{R_\mathrm{m}}$$
(6)
where $`V_{\mathrm{c},\mathrm{disk}}`$ and $`V_{\mathrm{c},\mathrm{halo}}`$ are the circular velocities due to the disk and halo respectively. In Paper I, we defined $`\eta _{\mathrm{exp}}`$ at $`R_{\mathrm{m},\mathrm{exp}}`$, the disk-plane radius at which $`V_{\mathrm{c},\mathrm{disk}}`$ would peak for an infinite, razor thin exponential disk, as appropriate for the model with the extensive halo; we then adopted $`R_\mathrm{m}=1.75R_\mathrm{d}`$ as the nearest equivalent for the Kuzโmin-Toomre disks. Here, however, we define $`\eta `$ at the true disk maximum for the Kuzโmin-Toomre disk, $`R_\mathrm{m}=1.41R_\mathrm{d}`$, since we employ that disk in all the simulations reported here. It should be noted that even though the rotation curve evolves as our simulations proceed, our values for $`\eta `$ are those of the initial model only.
Here we explore a wider range of models than in Paper I, with a greater variety of rotation curve shapes and find, not surprisingly, that a parameter which depends on the ratio of rotation velocities from disk and halo at a single radius does not correlate well with the degree of braking we observe in our simulations. Furthermore, halo angular momentum changes the evolution of $``$, so that $`\eta `$ is clearly an inadequate predictor of $``$ even for fixed rotation curves.
A parameter which takes into account both the angular momentum in the halo and the full rotation curve might do a better job. We define
$$\mathrm{\Gamma }(r_0)\frac{_{i,r<r_0}|J_{z,\mathrm{h},i}|J_{z,\mathrm{h},i}}{_{i,R<4R_\mathrm{d}}J_{z,\mathrm{d},i}}$$
(7)
where $`J_{z,\mathrm{h},i}`$ is the angular momentum of the $`i`$-th halo particle about the symmetry axis, $`r_0`$ is some arbitrary (spherical) cutoff radius for the summation and $`J_{z,\mathrm{d},i}`$ is the angular momentum of the $`i`$-th disk particle. The quantity $`\mathrm{\Gamma }(r_0)`$ is the difference between the maximum possible and the actual angular momentum content of the halo, expressed as a fraction of diskโs angular momentum, and is zero for a maximally rotating halo. It can be thought of as a measure of the capacity of the inner halo to accept angular momentum; it depends both on the halo mass distribution as well as its angular momentum content.
## 3 Massive Halo Models
We begin by describing a set of experiments with disks embedded in moderately dense halos. Halos of significantly greater density than we use here would inhibit the formation of the bar (Ostriker & Peebles 1973; Toomre 1981). We already demonstrated in Paper I that such a halo having an isotropic velocity distribution would brake the bar to an unacceptable extent. Here we determine the extent to which braking is affected by giving the halo net angular momentum, or an anisotropic distribution of velocities, or both.
### 3.1 Canonical simulation
Our canonical simulation (run 4) is the most halo-dominated model reported in Paper I, which we describe in more detail here. The initial rotation curve for this sub-maximal disk model (Figure 1a) drops unrealistically beyond the disk edge because the halo density drops smoothly to zero at the grid edge. The initial properties, numerical parameters and principal result are given in Tables 1 & 2.
The disk quickly forms a bar, as shown in Figure 2(a). Shortly after its formation, at $`t150`$, the bar buckled (Combes & Sanders 1981; Raha et al. 1991) very mildly, causing it to thicken. The continuing slow rise in $`๐`$ after this time results from gradual trapping of additional disk particles into the bar, often associated with spiral activity (Sellwood 1981); we describe this process as secondary bar growth. Spiral activity gradually declines as the outer disk heats to $`Q\mathrm{}>4`$. Figure 3 shows contour plots of the disk at several instances which clearly reveal the thickening of the disk, and the peanut shape of the bar. Figure 4 shows the distribution of disk particles at $`t=1000`$; note the prominent inner ring and the oval outer disk.
Once the bar forms, $`\mathrm{\Omega }_\mathrm{p}`$ has a well-defined value (Figure 2c) which drops rapidly at first, but reaches a constant level towards the simulationโs end at $`t=2000`$. Figure 2(d) shows that $`a_B`$ increases only mildly, whereas $`D_L`$ increases rapidly at first, later reaching a constant value, reflecting the behavior of $`\mathrm{\Omega }_\mathrm{p}`$ in Figure 2(c).
Figure 2(b) shows that the drop in $`\mathrm{\Omega }_\mathrm{p}`$ is associated with a substantial transfer of angular momentum from the disk to the halo. The torque which produces this angular momentum exchange arises from dynamical friction on the bar as it moves through the halo. The bar induces an $`m=2`$ response in the halo which develops very soon after the bar forms (Figure 5a). The response lags the position angle of the bar (Figure 5b) at first, but gradually shifts into alignment with the bar as the torque drops.
It is interesting that the bar survived such strong braking (cf. Kormendy 1979), which reduced its pattern speed by a factor of $`5`$. Most theoretical work, starting from Contopoulos (1980, see Sellwood & Wilkinson 1993 for a review) has suggested that self-consistent bars should nearly fill their corotation circles; our simulation is a clear counter-example. Its pronounced butterfly-shape when seen from above may be consistent with the prediction by Teuben & Sanders (1985) that slow bars require a large fraction of โboxโ orbits.
The rearrangement of angular momentum altered the density distributions in both the disk and halo, causing the rotation curve to change to that shown in Figure 1(b). The central density of the disk rose significantly but the density profile of the halo barely changed, despite all the work done on it by the bar (Figure 2b). Such a small change underscores how difficult it is for any dynamical interaction with the disk to modify the halo density profile.
We emphasize that a more realistic flat rotation curve model would require a more massive and extended halo. Not only might this increase dynamical friction, and slow the bar still more, but also the Lagrange point would lie further out in the disk, increasing the value of $``$; this effect alone would increase $`D_L`$ by more than 30% in this run at $`t=2000`$.
### 3.2 Halo rotation
All our models reported in Paper I, including our canonical model, have isotropic halos with no initial net angular momentum. Here we turn our attention to the effects of halos with net rotation. At first, we simply change the sign of $`J_z`$ for some halo particles according to the rule described in Appendix C.
We ran a series of experiments, summarized in Table 2, in which the halo angular momentum, $`=L_{z,\mathrm{h}}/L_{z,\mathrm{max}}`$, was $`\pm 0.33,\pm 0.67`$ and $`\pm 0.98`$, where $`L_{z,\mathrm{max}}`$ is the maximum achievable if the angular momentum of every particle is made positive. (The corresponding values of the dimensionless spin parameter, $`\lambda L|E|^{1/2}/(GM^{5/2})`$ are 0.05, 0.11 and 0.16, respectively). The mean rotation speeds of the halo particles are shown in Figure 6. Note that we did not continue all these simulations until the bar had finished slowing down.
Figure 7 shows that the value of $``$ reached by $`t=1000`$ correlates strongly with the angular momentum content of the halo. As found previously by Athanassoula (1996), the bars in models with retrograde halos are more strongly braked, while those in prograde halos less so, in comparison with the non-rotating case. Nevertheless, even when direct rotation in the halo is maximized, $`=1.7\pm 0.4`$ by the end of the run despite the fact that bar was weaker. The value of $`๐`$ in the maximally rotating models settled at some 70% of its value in the non-rotating simulation, suggesting that secondary bar growth, with a concomitant increase in friction, may be enhanced by angular momentum loss to the halo.
### 3.3 Anisotropic halos
We next present simulations with halos having somewhat more general DFs that yield anisotropic velocity distributions even in the absence of net rotation. Inspired by Osipkov (1979) and Merritt (1985), we chose the form $`(E+\beta J_z^2)`$, where $``$ has the same form as in equation (2). The halos are oblate and azimuthally biased when $`\beta <0`$ and prolate and radially biased when $`\beta >0`$. Anisotropy changes the distribution of mass within the halo; we therefore adjusted the halo mass fraction, $`f_\mathrm{h}`$, and/or the halo truncation radius, $`r_{\mathrm{halo}}`$, to generate models with values of $`\eta `$ similar to that in our canonical simulation.
Some initial tests revealed that too pronounced an azimuthal bias ($`\beta 0.2`$) led to strongly lop-sided halos which interfered considerably with the development of the bar. These $`m=1`$ instabilities, which appear to be of the kind discussed by Sellwood & Valluri (1997) for counter-rotating oblate spheroids, produced much larger asymmetries than those generally observed in real galaxies. We therefore report only those simulations which did not become strongly lop-sided. Disk-halo interactions via such $`m=1`$ modes are interesting in their own right and deserve a separate study.
The possible parameter space when the halos are allowed to be anisotropic is very large; we considered only two main models, an azimuthally biased model near the limit of $`m=1`$ stability, and a radially biased model (Table 2). Their rotation curves are shown in Figure 8. In both cases, we also tested fully-rotating versions of these models, which should have the weakest friction (cf. ยง3.2).
The bias towards large angular momenta introduced by setting $`\beta <0`$ tends to place more halo material at large radii at the expense of small radii and the inner rotation curve becomes dominated by the disk. To compensate, we further decreased the radial extent of the azimuthally biased halos. The oblate halo has an axis ratio $`0.7`$ at $`6R_\mathrm{d}`$. The bar that formed in a non-rotating halo was strongly braked, but friction is greatly reduced in the run with maximum halo rotation ($`\lambda =0.23`$).
The halos generated by radially biased DFs have larger densities at smaller radii, all other things being equal, than do isotropic halos. We therefore reduced the halo mass fraction in order to avoid a system which was too halo dominated. The halo was prolate at large radii, having an axis ratio $`1.16`$ at $`4R_\mathrm{d}`$, but it becomes oblate at $`R_\mathrm{d}`$ because of the diskโs influence. Since a trial run showed that secondary bar growth can be quite rapid in this model, we extended the disk to $`R_t=8R_\mathrm{d}`$. Once again, the bar is strongly braked in a simulation with no net halo rotation, but remains fast when the halo rotates maximally ($`\lambda =0.13`$).
### 3.4 Radially varying anisotropy and rotation
Tremaine & Ostriker (1999) suggest that the stringent limit on the halo central density that we obtained for isotropic halos (Paper I) could be relaxed if the halo had significant rotation in its inner parts only. They proposed that the inner halo had itself been torqued up and flattened by interactions with the disk.
To test their hypothesis, we create halos with varying anisotropy. We use polytrope-like distribution functions as before, $`(E+\beta J_z^2)`$, but now we let $`\beta `$ itself be a function of energy, $`\beta (E)`$:
$$\beta (E)=\beta _0(1\epsilon ^2)$$
(8)
where $`\epsilon =(EE_{\mathrm{min}})/(E_{\mathrm{max}}E_{\mathrm{min}})`$. Here, $`E_{\mathrm{max}}`$ and $`E_{\mathrm{min}}`$ are the values of the potential at the grid edge in the disk plane and the center of the system respectively. We set the free parameter $`\beta _0=0.2`$ in order to generate azimuthally biased, oblate halos, in line with the prediction of Tremaine & Ostriker. Figure 9 shows the axis ratio of the halo as a function of radius, which varies from $`0.5`$ at the center to spherical at the edge.
We ran a set of three experiments (runs 137-139 listed in Table 2) in which we introduced rotation by flipping retrograde halo particles with a probability that was a function of energy given by equation (C2), which gives something close to maximal rotation in the halo to some limiting energy; beyond some spherical radius $`r_{\mathrm{rot}}`$ the halo is non-rotating (Figure 9b). We varied this critical energy in this set of experiments. Two other experiments (runs 123 & 124) bracket them in terms of angular momentum content by having, respectively, no halo angular momentum and the maximum possible.
Again, the results are given in Table 2. As the angular momentum content of the halo is increased, the bar slows down less, but even when $`r_{\mathrm{rot}}=6`$ (Figure 9b) the final value of $``$ is still quite large. The parameter $``$ remains acceptably small only for the fully rotating case.
Mildly rotating halos in a similar set of experiments with a still more flattened halo ($`\beta _0=0.5`$), produced much weaker bars at first, because the bar buckled more violently; it seems that a different kind of coupling to the halo occurred in these cases. As these weak bars were slowed to a lesser extent after a fixed amount of evolution, we continued one simulation to $`t=3500`$, and found that the bar recovered and substantial friction again developed leading to a slow bar ($`=2.26\pm 0.20`$).
The mean orbital speed of the halo particles (Figure 9b) already indicates that significant halo rotation out to quite a large radius is needed to avoid strong friction. Figure 10 underscores just how much halo angular momentum this implies: we see that fast bars require $`\mathrm{\Gamma }(3)\mathrm{}<0.15`$ which is $`0.5`$ less than the value for the non-rotating halo. Thus the halo angular momentum inside $`r=3`$ has to be fully 50% of that of the entire disk.
Tremaine & Ostriker argue that strong halo rotation could be induced out to $`3`$kpc in a Hubble time, but we have shown here that the halo angular momentum requirements to avoid strong friction are considerably greater than their mechanism seems able to produce.
## 4 Halo Density and Concentration
In this section we report experiments in which the halo mass and concentration are varied, still with the halos confined to small radii, as in ยง3. In ยง5, we discuss more realistic models with extended halos. We have varied both the halo mass fraction and the polytrope index, $`n`$. Increasing $`n`$ leads to more concentrated halos, resulting in a larger halo contribution to the inner rotation curve (Figure 11). Table 2 lists the parameters of these simulations and gives our principal result. Note that these simulations represent three series, with varying halo mass density at fixed $`n`$.
The evolution of the $`n=3`$ runs has already been presented in Paper I. We find that decreasing the halo contribution to the rotation curve leads to a marked decrease in $``$. A fast bar (with a pattern speed that continued to decrease very slowly) survived in the simulation with the least massive halo of this series.
Figure 12 summarizes the values of $``$ obtained from models with different polytrope indices. Although there is no trend when all the points are taken together, within each series of runs at fixed $`n`$, bars with larger $`\eta `$ end up faster. Note that it would be incorrect to conclude from this Figure that less braking occurs for a given $`\eta `$ as the central density increases. This apparent trend results from two different effects: First, increasing $`n`$ for fixed halo mass leads to greater halo concentration, depleting halo material at larger radii in our halos (which we confined to a small volume), thereby reducing friction somewhat. At the same time, a more sharply peaked halo rotation curve leads to a smaller $`\eta `$ at fixed $`M_\mathrm{h}`$, and also makes the rotation curve drop more steeply, reducing $`D_L`$.
## 5 Models with more Extensive Halos
We next describe two models in which the halo extended to $`r=25R_d`$, which is large enough to achieve a flat rotation curve with moderate to low central densities in the halo. We had to increase the grid to $`257^3`$ โ an eight-fold increase over that used in most of the above experiments, making these runs much more expensive. A large parameter space search, such as that described in the previous two sections, would be prohibitively expensive with this larger grid.
We already reported a maximum disk model with an extensive halo in Paper I whose evolution was computed using a cylindrical polar grid. One of the two models presented in this section closely resembles it, but has a different disk type and is run on a Cartesian grid. The other model discussed here is a โcontrolโ experiment with a similar rotation curve but with a somewhat denser halo.
The disk, which was truncated at $`8R_d`$, accounts for 17% of the total mass. The polytrope index $`n=2`$ for the maximum disk model, whereas $`n=3`$ for the control run, this being the only difference between the two simulations. The resulting rotation curves, shown in Figure 13, are quite flat out to the disk truncation radius. Both have a substantial disk contribution at small radii: $`\eta =7.0`$ in the maximum disk case and $`\eta =3.8`$ in the control experiment.
The bars grew very rapidly and did not buckle much, reaching an amplitude some $`5`$% lower than in our canonical run. Most of the angular momentum lost by the inner disk after the bar forms ends up in the halo for both runs, but the outer disk continues to accept some of the barโs angular momentum.
Figure 14 shows the evolution of $`\mathrm{\Omega }_\mathrm{p}`$, $`a_B`$ and $`D_L`$ for both simulations. The bar in the control simulation is slow already by $`t1500`$ and $`=1.98\pm 0.35`$ by the end. The bar in the maximum disk simulation, on the other hand, is acceptably fast at the end of the run, with $`=1.57\pm 0.27`$, but only barely so and it is continuing to slow. We have therefore identified a region of parameter space where fast bars can survive for more than $`30`$ orbital periods at $`R=1.4`$ (just outside the half-mass radius of the disk).
## 6 Synthesis
We now seek some way to synthesize all these numerical results. We first show that the frictional torque acting on the bar from the halo behaves in some sense as might be expected from standard dynamical friction. However, the total angular momentum loss which occurs before the halo response locks into phase with the bar can be determined only numerically; we attempt to relate the final pattern speed of the bar to the ability of the halo to accept angular momentum.
### 6.1 The Chandrasekhar formula
Chandrasekharโs (1943) demonstration of a secular drag force on a massive perturber moving in a straight line through a uniform, infinite background sea of low mass particles differs in many significant ways from the rotation of a bar through an inhomogeneous halo. It is now part of received wisdom that his formula works better for a perturber moving through an inhomogeneous system, using local values of the density etc., than we have any right to expect (Binney & Tremaine 1987, ยง7.1). We here show that the frictional torque also behaves very roughly in accordance with his formula for a rotating bar โ at least over the period after the bar has formed and settled and before the halo response becomes aligned with the bar.
In the limit when the perturberโs mass $`M`$ is much larger than the masses of the background particles, Chandrasekharโs formula for the frictional force is
$$M\frac{d\mathrm{v}\mathrm{v}\mathrm{v}}{dt}=\widehat{\mathrm{v}\mathrm{v}\mathrm{v}}\frac{4\pi \mathrm{ln}\mathrm{\Lambda }G^2\rho M^2}{\sigma ^2}g\left(\frac{v}{\sigma }\right),$$
(9)
where $`\mathrm{v}`$$`\mathrm{v}`$$`\mathrm{v}`$ is the velocity of the perturber, $`\widehat{\mathrm{v}\mathrm{v}\mathrm{v}}`$ is a unit vector in the same direction and $`\rho `$ and $`\sigma `$ are respectively the density and velocity dispersion of the background. The $`v^2`$ factor in the denominator of formula (7-18) of Binney & Tremaine has here been replaced by $`\sigma ^2`$ in order to subsume all the velocity dependent factors into the dimensionless function $`g`$, which is shown in the Figure 15(a) for the case when the velocity distribution of the background particles is Gaussian. It can be seen that friction increases as the speed of the perturber rises, reaching a maximum when $`v1.37\sigma `$ and then decays monotonically for all higher speeds.
The rate of gain of angular momentum of the halo in our simulations is clearly the frictional torque $`\tau _z`$ on it arising from the bar. Note that this measurement is independent of barโs reaction to the loss of angular momentum, and therefore does not involve, for example, any estimate of its effective moment of inertia. The halo torque could be interpreted as the frictional force times some characteristic lever arm of length $`R`$. We therefore plot in Figure 15(b) the quantity
$$T_z=\frac{\tau _z\sigma ^2}{R๐^2\rho }$$
(10)
against the speed of the bar perturbation through the background halo at that radius, $`v^{}=R\mathrm{\Omega }_\mathrm{p}v`$, normalized by the halo velocity dispersion. The running averages in this expression are over 50 time units, and we include $`๐`$ to take account of the variations in bar mass in these equal mass disks. The scaling of the ordinate is arbitrary, therefore. We adopt $`R=3R_{m=2,\mathrm{max}}`$ (i.e. three times the radius where the $`m=2`$ Fourier component peaks); other values of $`R`$ show the same general trend, but we found that this choice minimized the scatter. We adopt local values for $`\rho `$ and $`\sigma `$ (averaged over the range $`0<R<3R_{m=2,\mathrm{max}}`$), and we use $`\sigma ^2=\sigma _R^2+\sigma _\varphi ^2+\sigma _z^2`$.
We plot a curve from one simulation in Figure 15(b) and a number of points from other simulations. The curve shows the entire evolution of $`T_z`$, from the moment the bar first forms to the end when friction is very low, for a maximally retrograde model (run 20). The time evolution along this line is from right to left and can be described as follows: (1) the initial spike occurs as the halo response develops soon after the bar forms, (2) the curve then dips as the bar buckles, (3) after which the line follows roughly the trend indicated by the points as the bar is braked steadily, and (4) it drops down to low values as the halo response locks into alignment with the bar. The fluctuations in $`T_z`$ in this run are fairly typical; they are larger in some models and less in others.
We should not expect Chandrasekhar-style friction in any part of this evolution except for period (3) after a quasi-steady halo response is established, and before orbit trapping becomes significant. Thus we have tried to include points in this plot from other runs during the steady friction period, although we obviously failed for the cluster of points near $`v^{}/\sigma =0.6`$ and $`T_z\mathrm{}>0`$.
Although there is considerable scatter in these measurements, the general rise over the range of the abscissae is similar to the theoretical curve in Figure 15(a), although our data show only a rising trend. We are encouraged that a trend shows through despite the crude approximations we adopt: we identify a single radius whereas the entire halo probably contributes, not all our bars have identically the same density profile, our polytrope halos do not have a precisely Gaussian velocity distribution, etc. While perhaps not entirely convincing, this Figure does suggest (i) that the torque is indeed from the usual dynamical friction, (ii) that most of the drag seems to arise from the region just beyond the end of the bar, and (iii) that the characteristic speed is generally less than the halo dispersion.
It should be noted that the velocity dependence in Figure 15(b) is the opposite of that predicted by Weinberg (1985), who suggested that friction would increase if the halo was made to rotate in a prograde sense; one interpretation of Weinbergโs prediction is that in his case most of the friction arises from a perturber speed $`v\mathrm{}>1.4\sigma `$. The fact that we find the opposite behavior may result from his assumption of an infinite isothermal sphere for the halo, whereas our halos have a very limited radial extent.
### 6.2 Constraints from $`\mathrm{\Gamma }`$
In our simulations with strongly confined halos, friction seems to drop to zero before the bar is brought to rest relative to the streaming speed of the halo particles. The torque vanishes when the induced bi-symmetric distortion in the halo locks into alignment with the bar. This locking effect appears to be the result of non-linear trapping of orbits, a process described as โdynamical feedbackโ by Tremaine & Weinberg (1984b). Note that this locking phenomenon did not occur in our simulations with more extended halos, where braking persisted for as long as we ran them (see Figure 14).
We have searched for a predictor of the final bar pattern speed, but have not found anything simple โ perhaps because none exists. The best we have come up with is the parameter $`\mathrm{\Gamma }`$ introduced in ยง2.3. After some experimentation, we found $`\mathrm{\Gamma }(3)`$, evaluated using the initial values of $`J_{z,h}`$ and $`J_{z,d}`$, to be the most useful. Figure 16 summarizes the last measured values of $``$ from an ensemble of 43 different simulations plotted against $`\mathrm{\Gamma }(3)`$. These simulations include various different polytrope indices, different anisotropies, different halo rotations (and different distributions of halo angular momentum), different Toomre $`Q`$s, different disk thicknesses and different halo masses. Not all of these simulations have been evolved to a steady state; we generally stopped the simulation once we found the bar to be slow (which may account for some of the scatter in the distributions).
We draw the following conclusions from the rising trend in this Figure: (1) Bars can generally remain fast when $`\mathrm{\Gamma }(3)<0.4`$, although some are significantly braked. (2) When $`\mathrm{\Gamma }(3)\mathrm{}>0.4`$, bars generally end up slow. Strong braking can be avoided when $`\mathrm{\Gamma }(3)\mathrm{}>0.4`$ when either the model has a rapidly dropping rotation curve, or a flattened halo with most of its mass outside the bar region.<sup>2</sup><sup>2</sup>2The positions of models with radially varying anisotropy in this plot is more than usually sensitive to the value of $`r_0`$ used in computing $`\mathrm{\Gamma }`$.
Our parameter $`\mathrm{\Gamma }`$ is the best we have been able to find to predict the value of $``$. A range of values of $`\mathrm{\Gamma }`$ can be determined for any rotation curve fit, but, being dark, no value of $`\mathrm{\Gamma }`$ can be pinned on a halo. Constraints on the actual angular momentum content of dark halos might be possible from studies of the faint luminous halos that may trace the rotation of the dark halo (being subject to similar dynamics). The Tremaine-Ostriker hypothesis calls for low $`\mathrm{\Gamma }`$ due to high rotation in the halo. We favor low $`\mathrm{\Gamma }`$ due to low halo mass fraction (we have argued that the Tremaine-Ostriker hypothesis may need rather unlikely levels of angular momentum in the inner parts of halos). The reality may be somewhere in between these two limits.
## 7 Discussion
### 7.1 Comparison With Real Bars: NGC 936
We need to show that our $`N`$-body bars are similar in strength to bars in real galaxies. A photometric comparison would not be conclusive because the strongly non-axisymmetric light distribution in a galaxy may not reflect the true distribution of mass. We have therefore attempted a kinematic comparison using data from NGC 936, a well-studied SB0 galaxy having sufficient published data on the stellar velocity field for our purposes. Its other advantages are that it appears to be relatively dust-free and is known to have a fast bar (Kent 1987; Merrifield & Kuijken 1995). The galaxy has a dense bulge (Kent & Glaudell 1989), however, unlike in our simulations.
We adopt Kormendyโs (1983) estimates of the projection geometry, the bar position angle, the de-projected bar semi-major axis and use his velocity measurements at five different slit position angles. We scale our models by setting $`a_B=50^{\prime \prime }`$ and rotate and project them as we view NGC 936. We then compute the mean projected line-of-sight velocity ($`V_{\mathrm{los}}`$) of the particles in the model, averaging the approaching and leading sides to maximize the number of particles in our pseudo-slits, and determine the velocity scaling by minimizing $`\chi ^2`$ between the observations and the model. When making this comparison, we use data in the circular annulus (in the galaxy plane) $`0.6R/a_B1.2`$ to exclude the region dominated by the bulge and the disk well outside the bar.
Our canonical model at early times compares very well with NGC 936. We find reduced $`\chi ^2=0.70`$ (for 22 degrees of freedom) at $`t=250`$, and a visual comparison of $`V_{\mathrm{los}}`$ shows that the variations with position angle track those in the galaxy very well, as they must do for this good a fit. (For comparison, we obtain a reduced $`\chi ^2=1.96`$ if we erase the bar from our model by randomizing the azimuthal positions of the particles.) At later times, after the bar has been slowed by a factor $`5`$, we find reduced $`\chi ^2=5.94`$, again showing that NGC 936 is grossly inconsistent with a slow bar.
The maximum disk model is not quite as impressively similar to NGC 936: reduced $`\chi ^2=1.20`$ at $`t=350`$ which is still acceptable, but reduced $`\chi ^2=1.54`$ at $`t=1150`$, which is marginally so. The fits to โcontrolโ run are again worse: reduced $`\chi ^2=1.61`$ at $`t=318`$, when the bar is still fast, and reduced $`\chi ^2=1.75`$ at $`t=2000`$ when the bar has slowed. In both cases, the absence of a massive central spheroidal bulge may be partially responsible for the poorer fits.
Thus the bars in our simulations are quite similar to that in NGC 936 when they first form, but are clearly inconsistent when they have been strongly braked.
### 7.2 Neglected Physics
Our simulations are of the stellar and DM components of a barred galaxy and do not include any gas. It is well known that gas flows in barred potentials produce large-scale shocks offset from the bar major axis (e.g. Athanassoula 1992). The asymmetric gas distribution loses angular momentum to the bar and gas flows towards the center. The small gas fraction in most galaxies, together with the relatively short lever arm, mean that the angular momentum given to the bar by gas could not possibly compete with that lost to the halo through dynamical friction โ e.g. friction removed $`40`$% of the diskโs angular momentum in our canonical simulation (Figure 2). Furthermore, gas-poor SB0 galaxies, such as NGC 936 and NGC 4596, have fast bars.
Gas inflow has a second effect, however: it deepens the gravitational potential at the bar center causing the bar to speed up slightly. This can be a small effect at most, since excessive inflow will destroy the bar. The total mass accumulated in the center cannot exceed 5% of the disk mass (Norman et al. 1996), and is probably much less; even in this extreme case, the increase in bar pattern speed was only some 40% (Sellwood & Debattista 1996, Figure 6).
A bulge component inside the inner Lindblad resonance might act as a source of angular momentum for the bar. Bulges can be quite massive (e.g. NGC 936, Kent & Glaudell 1989) and are often in rapid rotation (e.g. Kormendy & Illingworth 1982), but their small radial extent limits their angular momentum content. It is possible to think of the inner parts of the rapidly rotating halo in some of our simulations as representing a bulge. The bar is still strongly braked in such cases (ยง3.2), suggesting that not even rapidly rotating bulges can prevent bar slow-down in sub-maximal disks. Weinberg (1985) suggested that the inner disk could also be a source of angular momentum for the bar, but we have found that bars always lose angular momentum to the disk, not the other way round.
We have not included the effects of late gas infall onto the disks which would enhance secondary bar growth (e.g. Sellwood & Moore 1999). Irrespective of the rate of bar growth by this mechanism, it inevitably leads to increased friction, making it still more difficult for the value of $``$ to decrease.
In addition to massive gas inflow, bars can be destroyed by satellite impacts. Our simulations do not take either process into account. If these processes are to account for the absence of slow bars, they would have to act efficiently and quickly, and new bars would have to form again to maintain the observed high fraction of barred galaxies. The rapid formation of a new bar is difficult to arrange, since bar destruction leaves the disk dynamically hot, and the inflow destruction mechanism gives rise to a more steeply rising inner rotation curve. Both factors limit the diskโs ability to form new bars, by making it less responsive and by cutting the feed-back loop to the swing amplifier (Toomre 1981). The revival of a bar after one has dissolved requires the accretion of so much dynamically cool material (Sellwood & Moore 1999) that it is unlikely to occur more than once in a galaxy.
While the bars in our simulations are comparably strong to that in a real galaxy (ยง7.1), a systematic difference with early-type galaxies is the absence of a dense bulge in our simulations. It is possible that simulations with dense bulges behave differently, but it seems unlikely that they would. Orbit studies (Athanassoula 1992) and simulations (Sellwood 1989; Sellwood & Moore 1999) reveal that bars in galaxies with dense centers are supported by the same orbit family and behave similarly to those in which the center is more uniform.
### 7.3 Scaling to NGC 3198
Before discussing the implications of our results for real galaxies, we need to determine how they should be scaled. The de facto standard galaxy in the dark matter halo debate is NGC 3198; even though it is not a strongly barred galaxy, we nevertheless scale our models to the data of van Albada et al. (1985) for this system.
We use our two extensive halo systems from ยง5 for this comparison. Scaling to the observed rotation curve determines the length and the velocity scales. We first adopted $`R_d=3.0`$ kpc for both simulations, which differs from the exponential scale of 2.68 kpc that fits NGC 3198 well (van Albada et al. 1985) since our models have Kuzโmin-Toomre disks. We then adjusted the velocity scale to minimize the residuals between the data and our model rotation curves, finding $`(GM/R_\mathrm{d})^{1/2}=584`$ & 540 km s<sup>-1</sup> for the maximum disk and control runs, respectively. The resulting scaled rotation curves are shown in Figure 17, which also shows the maximum-disk fit of van Albada et al. As usual, both models fit NGC 3198 very well, which is another instance of the rotation curve degeneracy. Note that both systems are quite disk dominated, and that the maximum disk of van Albada et al. is perhaps even more disk dominated than our โmaximum diskโ simulation.
Choosing length and velocity scales determines the time unit. With these adopted values, the duration of many of our experiments, $`\mathrm{2\hspace{0.17em}000}`$ dynamical times, is equivalent to $`10`$ Gyr.
## 8 Conclusions
### 8.1 Summary of Principal Results
We have shown that the severe braking of the bar by dense isotropic halos reported in Paper I also occurs for other non-rotating, or backwards rotating, halos of similar density, whatever the shape of the halo velocity ellipsoid. In all such cases, the bar in the disk slows unacceptably in a few rotations. The bars in all our experiments with strongly prograde rotation in the halo were not braked as severely; the halo spins strongly in those cases for which friction ceased before $``$ rose above 1.4.
The existence of strong friction is in agreement with the theoretical prediction by Weinberg (1985), with earlier fully self-consistent simulations by Sellwood (1980), using a coarse grid, and by Athanassoula (1996) using a different $`N`$-body method, and other work (e.g. Hernquist & Weinberg 1992). We also find that all our bars slow down as they lose angular momentum โ a non-trivial result since bars are not rigid objects and could conceivably spin up (e.g. as a binary star) as angular momentum is lost. Remarkably, the bars in some of our experiments survived a truly drastic slow-down โ more than a factor of five in many cases โ providing further evidence that bars are in fact dynamically very robust objects (Miller & Smith 1979; Sellwood & Wilkinson 1993). We find no evidence to support the idea (Kormendy 1979) that strong braking of a bar might cause it to dissolve.
In some of our simulations, the bar did a great deal of work on the halo โ e.g. in the canonical run the halo gained 40% of the angular momentum of the disk. Nevertheless, the change in the halo density profile was quite minor (Figure 1b). This example emphasizes that it is extremely difficult to change the density profile of a halo using interactions with the disk.
Friction is generally reduced by lowering the density of the halo, and bars in maximum disks are able to remain fast (though only barely so) for large numbers of dynamical times, even in extensive, non-rotating halos, as reported in Paper I. The bar in our maximum disk simulation with an extensive halo is continuing to slow down even after 2000 dynamical times (Figure 14a), suggesting that $``$ might continue to rise. When scaled to NGC 3198, this continued evolution is too slow to matter, since we have followed it for 10 Gyr. But dynamical times are shorter in more luminous galaxies and our computed evolution lasts the equivalent of 5 Gyr in a galaxy with $`V_{\mathrm{max}}300`$km s<sup>-1</sup>.
Friction is caused by a lag between the bar and an $`m=2`$ distortion in the halo (Figure 5). None of our halos is rotating sufficiently rapidly to be itself unstable to bar-forming modes (e.g. Sellwood & Valluri 1997), so the halo distortions are responses forced by the rotating bar in the disk, as is usually the case in dynamical friction. It is encouraging that we have been able to find some suggestion of the expected velocity dependence of the frictional force in our very crude analysis (Figure 15), which suggests that the qualitative effect of halo rotation is predictable.
Braking ceases once the forced response in the halo rotates in alignment with the bar in the disk. The gradual trapping of halo orbits into the driven non-axisymmetric potential is itself one of the principal sources of dynamical friction. It seems reasonable that trapping of halo orbits should involve less angular momentum loss for the bars in halos with an excess of particles with $`J_z>0`$, as we have found. Note that we have observed the locking of the halo response into alignment with the bar only in models with strongly confined halos.
### 8.2 Implications for barred galaxies
As noted in the introduction, the rather small number of actual measurements of real barred galaxies all lie in the range $`11.4`$; the existence and locations of dust lanes in bars is indirect evidence that these low values pertain more generally. Thus our results require either that most strong bars are (1) too young to have been slowed significantly, (2) exist in strongly rotating halos, or (3) are not embedded in dense halos. We review each of these in turn.
If disks are not maximal, and halos do not rotate strongly, then bars must indeed be young objects to have remained fast today. Since the rate of bar slow-down scales with the halo density, the larger the required density, the younger they must be to avoid any slow cases. There is a suggestion that bars were rare in the early universe (Abraham et al. 1999), but they have certainly existed, probably in about their present numbers, since a redshift of one half. Their ages are therefore $`\mathrm{}>4`$Gyr, or $`\mathrm{}>800`$ dynamical times when our simulations are scaled to NGC 3198, which is plenty long enough for friction to have slowed the bars significantly, although perhaps not completely.
Bars in galaxies which are significantly sub-maximal can remain fast for cosmologically interesting times only if the halo is anisotropic and rotates strongly throughout most of the disk region. The required halo angular momentum is very large, however. If all halo angular momentum arises from tidal torques in the early universe, the required value of $`\lambda `$ would be reached only rarely (Barnes & Efstathiou 1987; Steinmetz & Bartelmann 1995). Since $`\mathrm{}>50\%`$ of all HSB disks containing strong bars (Eskridge et al. 2000), the vast majority cannot have $`\lambda `$ large enough to avoid bar slow down.
Alternatively, one could imagine the inner halo to have been torqued up by some means. Tremaine & Ostriker (1999) suggest two ways to transfer angular momentum from the disk to the inner halo for precisely this purpose. We have found that if the halo has moderate central density, then it must have a high degree of rotation โ fully half that of the disk, out to well beyond the barโs end. If the nearest strongly barred galaxy, our own Milky Way, has a sub-maximal disk, we require substantial rotation in the halo within the Solar radius for the bar to have avoided strong braking. Most torquing mechanisms would act equally on both the dark halo and any associated stars. Thus the absence of significant rotation in, for example, the metal-weak globular cluster system of the Milky Way (Harris 2000) also suggests that the inner halo lacks the required angular momentum.
We therefore conclude that bars in real galaxies remain fast because disks are maximal. Weiner et al. (2000) reach a similar conclusion on quite different grounds in the case of the barred galaxy NGC 4123.
### 8.3 General Implications
It is often argued (e.g. Courteau & Rix 1999) that barred galaxies have lower density halos than do unbarred. This prejudice stems from the paper by Ostriker & Peebles (1973), who suggested that the only way to inhibit bar formation in a galaxy was to immerse the cool disk in a massive dynamically hot component. Not only is this argument fallacious (Toomre 1981; Sellwood & Evans 2000), but we here present evidence that bright barred galaxies have similar DM fractions as do their unbarred counterparts.
In Paper I, we argued against the hypothesis that barred galaxies have less DM than those of the unbarred family: If the DM content varies continuously between maximum disk, fast bar, SB galaxies and massive halo SA galaxies, there should be many galaxies of intermediate dark matter content. Any strong bars that may form in such galaxies would therefore be fiercely braked, as in our experiments. Since no slow bars are known in HSB galaxies, we conclude that, either the distribution of dark matter is bimodal, or that all galaxies with moderate halo density have somehow avoided forming bars, both of which seem very unlikely, or that no galaxies are dark matter dominated.
Tidal triggering can induce a bar in a galaxy that is stable when isolated (e.g. Noguchi 1987). Such bars would be strongly braked if the target galaxy had been stabilized by a massive halo. The absence of known slow bars again argues against massive halos, but only weakly; if the rate of bar-inducing tidal interactions is low, then the measured sample may be simply too small to include a slow case.
Empirical evidence against a systematic difference between barred and unbarred galaxies was presented by Bosma (1996) and more can be found in the data from Mathewson & Ford (1996). We use the apparent magnitudes in the I-band, recession velocities and $`V_m`$ given in their table, convert to absolute magnitudes assuming Hubble distances (for $`H_0=60`$km s<sup>-1</sup> Mpc<sup>-1</sup>) and plot the Tully-Fisher relation for the 2368 galaxies in their sample having recession speeds $`>1,000`$km s<sup>-1</sup> (to avoid absurdly inaccurate Hubble distances) in Figure 18(a). The line is fitted to all the data, but the 2219 points are for โunbarredโ galaxies and the crosses mark the 149 galaxies which Mathewson & Ford designate as barred.<sup>3</sup><sup>3</sup>3Their sample excluded galaxies designated as barred but, as always happens, bars were identified after the data were taken. It is unclear whether these are typical bars, but since their barred fraction is extremely low, it seems likely that they flagged only the blatant, i.e. strong, bars. The histograms in Figure 18(b) show the distributions of velocity residuals about the fitted line for the barred and unbarred galaxies separately (scaled so that the area under each histogram is unity), suggesting a small offset in the sense that the barred galaxies have slightly lower $`V_m`$ at a given brightness. A Kolmogorov-Smirnov test indicates that there is 3.5% probability that these two samples were drawn from the same parent distribution, suggesting a possibly significant difference. However, the offset disappears if we discard all galaxies fainter than M$`{}_{\mathrm{I}}{}^{}21`$, indicating that it arises from the faint galaxies only, as is apparent in Figure 18(a). We conclude that there is no evidence here for a deficiency in DM content, relative to the unbarred galaxies, in the (few) bright barred galaxies in the Mathewson & Ford sample. Further Tully-Fisher work with properly constructed samples of barred/unbarred galaxies to confirm this conclusion would be highly desirable.
We conclude that all bright HSB disk galaxies, barred or unbarred, are maximum disks. Supporting evidence for this conclusion is reviewed by Sellwood (1999).
We also predict that any barred galaxy having a moderately dense halo should have a slow bar. Prime candidates to test this prediction may be found amongst galaxies believed to have significant DM fractions in their inner regions: the low luminosity galaxies (e.g. Persic & Salucci 1988; see Sellwood 1999) and low surface brightness galaxies (LSBs, Bothun et al. 1997; de Blok & McGaugh 1997). Bars in these systems are less common, but not unknown.<sup>4</sup><sup>4</sup>4The slight offset between the fainter barred and unbarred galaxies in Figure 18(a) is in the sense that the barred cases have a lower $`V_m`$ in relation to their luminosity, and presumably therefore a smaller DM fraction, than do the unbarred. It is reasonable that bars should be found amongst those galaxies with more dominant disks. If a strongly barred low-luminosity or LSB galaxy has even a moderately dense DM halo, it should have a high value of $``$. Unfortunately, there are no reliable measurements of pattern speeds in such galaxies to test the prediction at this time.
We would like to thank Scott Tremaine for a number of insightful conversations and a critical read of the manuscript. The comments of the referee, Lia Athanassoula, were also helpful. This work was supported by NSF grant AST 96/17088 and NASA LTSA grant NAG 5-6037. VPD acknowledges support of grant # 20-50676.97 from the Swiss National Science Foundation for part of this work.
## Appendix A Determination of an Equilibrium Distribution Function
We here describe the creation of our equilibrium halo models in the presence of a massive disk. The procedure is identical to that followed by Raha et al. (1991), but has not previously been explained in detail in a published article. Because the equilibrium generated by this procedure is exact, there is no need to compute the evolution of the halo while it adjusts to an equilibrium from an approximate set-up (e.g. Barnes 1992; Hernquist 1993).
We adopt the iterative approach to finding a DF first proposed by Prendergast & Tomer (1970) and developed by Jarvis & Freeman (1985) for two component systems and Kuijken & Dubinski (1995) for a three component system. Unlike these authors, however, we solve for the gravitational potential using the same numerical procedure that is used in the simulations, thereby incorporating any numerical idiosyncrasies of the potential determination into the solution for the DF; this strategy ensures that the particle distribution is in perfect equilibrium at the outset.
We first choose a functional form for the DF of the halo
$$f=C(I_1,I_2,\mathrm{}),$$
(A1)
where $`C`$ is a normalization constant and $``$ can be more or less any reasonable function of the classical isolating integrals, $`I_n`$. In our axisymmetric potential, these are $`E`$ and $`J_z`$. The form of $``$ adopted determines the density profile and shape of the resulting halo; functions of $`E`$ alone tend to produce almost spherical halos (the disk makes them slightly oblate), while adding a dependence on $`J_z`$ generally produces more strongly spheroidal halos.
A first approximation to the halo density $`\rho _1(R,z)`$ can be determined from
$$\rho _1(R,z)=fd^3\mathrm{v}\mathrm{v}\mathrm{v}$$
(A2)
using any reasonable initial guess for the gravitational potential $`\mathrm{\Phi }_1(R,z)`$. We assign mass to the grid to represent the smooth function $`\rho _1(R,z)`$, add the mass distribution of a smooth disk and solve for a new gravitational field $`\mathrm{\Phi }_2(R,z)`$. We then determine $`\rho _2(R,z)`$ using the improved potential in (A2), and iterate until the potential distribution converges. The value of $`C`$ can be adjusted at each iteration step to drive the solution to the desired halo mass. We find that the solution converges rapidly and that 10 iterations are usually ample.
Note that although the halo density profile, and therefore the net rotation curve, cannot be specified in advance, the rapid convergence permits many models to be explored (for different mass ratio, truncation radius, choice of $``$, etc.) from which one having the desired properties may be selected.
## Appendix B Quiet Start Procedures
Since an $`N`$-body simulation amounts to a numerical solution of the coupled collisionless Boltzmann and Poisson equations by the method of characteristics, it is clearly desirable to select the characteristics to be followed with care. Selecting particles at random from a DF (e.g. Hernquist, Spergel & Heyl 1993; Kuijken & Dubinski 1995) leads to $`\sqrt{N}`$-type variations in the number of particles generated in any given range of the integrals; in effect the model will have the dynamical properties of one with a DF slightly different from that intended, which has many disadvantages, especially when attempting comparisons with theoretical work. The following quiet start procedures lead to much higher quality simulations (and are also more efficient); every part has been described in some other publication, but for ease of reference we summarize them here.
Strategies for the optimal selection of points are exactly analogous to those for the selection of abscissae in the numerical evaluation of multi-dimensional integrals. In that case, accuracy is improved whenever the dimensionality can be reduced by analytic integration over some of the coordinates. In our problem, we know the DF to be independent of orbital phases, since they must be uniformly populated in any equilibrium model. Note that, except when the DF is expressed in terms of actions, the density of particles in the sub-space of the integrals is not simply given by the DF; it needs to multiplied by a โdensity of statesโ function, which is the phase-space volume per unit interval of $`E`$ and $`J_z`$ (Binney & Tremaine 1987, ยง4.4.5)
For a razor-thin disk, the density of particles having energy $`E`$ and angular momentum $`J_z`$ is (Sellwood & Athanassoula 1986)
$$๐ฉ_{\mathrm{disk}}(E,J_z)=2\pi f(E,J_z)\tau (E,J_z),$$
(B1)
where $`f`$ is the usual phase space density and $`\tau (E,J_z)`$ is the period of one complete radial oscillation of a particle with the given $`(E,J_z)`$. The latter generally has to be determined numerically. For a sphere with $`f(E,L)`$
$$๐ฉ_{\mathrm{sphere}}(E,L)=8\pi ^2Lf(E,L)\tau (E,L),$$
(B2)
(Binney & Tremaine 1987, problem 4-22), while, for a spheroid with $`f(E,J_z)`$,
$$๐ฉ_{\mathrm{spheroid}}(E,J_z)=4\pi ^2f(E,J_z)S(E,J_z)$$
(B3)
(Sellwood & Valluri 1997). In this last formula, $`S(E,J_z)`$ is the cross-sectional area in $`(R,z)`$ of the bounding torus in the meridional plane (Binney & Tremaine 1987, ยง3.2.1) and is easily evaluated numerically for arbitrary potentials.
We proceed by slicing accessible $`(E,J_z)`$ space into a number, $`j=n_En_J`$, of small areas in such a way that the integral of the DF over each area encloses a fraction $`1/j`$ of the total active mass; we generally choose $`n_En_J`$. We then select a point within each of these areas to determine the $`(E,J_z)`$ values for an orbit. We avoid a regular raster of such points in $`(E,J_z)`$ space while maintaining the desired smoothness, as follows: For every slice in energy, we choose $`n_J`$ equal spaced values of $`J_z`$ from the distribution of $`๐ฉ|_E`$, with the first value only determined as a random sub-fraction, and then select an $`E`$ value within each area at random from the distribution of $`๐ฉ(E,J_z)๐J_z`$.
Each $`(E,J_z)`$ pair selected in this way specifies an orbit and we must next choose phases to determine both the initial position and velocity components of the particles. In contrast to the selection of integrals, experience suggests that the behavior of the model is much less sensitive to the manner in which some orbital phases are selected. In general, random selection from the appropriate distribution is adequate, although it is easy to improve upon random when desired for a particular application. Two examples are: Sellwood (1983) found it desirable to space several otherwise identical particles equally around a ring when searching for small-amplitude non-axisymmetric instabilities and Sellwood (1997) was able to quieten radial pulsations of a stable spherical model by spacing particles equally in radial phase.
In a razor-thin disk, or in a sphere, the orbit lies in a plane in which particles oscillate between peri- and apocenter with full period $`\tau (E,L)`$. The radial phases must be uniformly distributed, but the probability of selecting a particular radius varies inversely with the (non-uniform) radial speed. It is easiest to select a fraction of the radial period and integrate the orbit (usually numerically) for this time to determine the radius. The radial and azimuthal speeds are completely determined (except for the sign ambiguity of the radial speed) by the selected values of $`E`$, $`J_z`$ and $`r`$. The azimuthal phase and, for a sphere, the orientation of the plane can be selected in a straightforward manner.
In spheroidal models, as here, the two classical integrals confine the particle to a torus in real space. When the desired DF does not depend upon a third integral, the probability density distribution for any given orbit is uniform in the meridional plane within the boundaries of the torus and selection of $`(R,z)`$ pairs is straightforward. The values of $`E`$ and $`J_z`$ almost determine the velocity components at the chosen position โ all that remains is to direct the velocity component in the meridional plane, which should be uniformly distributed in angle.
## Appendix C Halo Angular Momentum
In a general axisymmetric system, the dependence of the density $`\rho `$ on $`J_z`$ is only through the even part of the DF (Lynden-Bell 1962). Thus angular momenta about the symmetry axis $`z`$ can be reversed at random without affecting the equilibrium of the system. Changing the net angular momentum may, however, alter the stability of the system; in particular, Kalnajs (1977) has cautioned that a discontinuity in the DF at $`J_z=0`$ can aggravate instabilities. We therefore adopt a scheme which ensures a smooth DF.
If a prograde halo is desired, we define $`p_\alpha (x)`$, where $`x=J_z/J_{z,max}`$, to be the probablity of changing the sign of $`J_z`$ of a retrograde particle in a halo in which the maximum possible angular momentum at the truncation energy is $`J_{z,max}`$. (To generate a retrograde halo, we flip only particles having positive $`J_z`$ with probablity $`p_\alpha (x)`$, where now $`x=J_z/J_{z,max}`$.)
In order to make the DF continuous at $`J_z=0`$, we require $`p_\alpha 0`$ as $`x0`$. By extending the definition of $`p_\alpha (x)`$ to be an odd function when $`x<0`$, we can write the new distribution function as $`\stackrel{~}{f}(E,J_z)=f(E)[1p_\alpha (x)]`$.
We adopted the shifted, normalized Fermi function and its inverse, which both have the desired properties:
$$p_\alpha (x)=\{\begin{array}{cc}\left[(e^{\alpha x}+1)^1\frac{1}{2}\right]/\left[(e^\alpha +1)^1\frac{1}{2}\right]\hfill & \mathrm{large}L_{z,h}\hfill \\ \frac{1}{\alpha }\mathrm{ln}\left(\left\{x\left[(e^\alpha +1)^1\frac{1}{2}\right]+\frac{1}{2}\right\}^11\right)\hfill & \mathrm{small}L_{z,h}\hfill \end{array}$$
(C1)
When $`\alpha =0`$, both functions are $`p_0(x)=x`$, which leads to a certain total $`L_{z,h}`$. If the desired total $`L_{z,h}`$ is greater (smaller) than this value, we use the large (small) expression and adjust $`\alpha `$ to yield the desired net angular momentum.
In order to generate a radial variation in the net rotation, we make the probability a function of energy as follows:
$$p_d(x)=N(E)\frac{\left(e^{\alpha x}+1\right)^1\frac{1}{2}}{\left(e^\alpha +1\right)^1\frac{1}{2}}$$
(C2)
with
$$N(E)=\{\begin{array}{cc}1\hfill & EE_1E_{\mathrm{min}}+(E_{\mathrm{max}}E_{\mathrm{min}})\frac{1}{d}\hfill \\ 2s^33s^2+1\hfill & E_0>E>E_1\hfill \\ 0\hfill & EE_0E_{\mathrm{min}}+(E_{\mathrm{max}}E_{\mathrm{min}})(\frac{1}{d}+\frac{1}{10})\hfill \end{array}$$
(C3)
Here $`s=(EE_1)/(E_0E_1)`$, while $`E_{\mathrm{min}}`$ and $`E_{\mathrm{max}}`$ are constants of the system. We fix $`\alpha =30`$ to ensure the inner halo rotates strongly and vary $`d`$ to adjust the energy at which strong rotation gives over to no rotation.
|
warning/0006/physics0006067.html
|
ar5iv
|
text
|
# 1 Problem
## 1 Problem
Deduce the velocity distribution of steady flow of an incompressible fluid of density $`\rho `$ and viscosity $`\eta `$ between two parallel, coaxial annular plates of inner radii $`r_1`$, outer radii $`r_2`$ and separation $`h`$ when pressure difference $`\mathrm{\Delta }P`$ is applied between the inner and outer radii.
As an exact solution of the Navier-Stokes equation appears to be difficult, it suffices to give an approximate solution assuming that the velocity is purely radial, $`๐ฏ=v(r,z)\widehat{๐ซ}`$, in a cylindrical coordinate system $`(r,\varphi ,z)`$ whose $`z`$ axis coincides with that of the two annuli. Deduce a condition for validity of the approximation.
This problem arises, for example, in considerations of a rotary joint between two sections of a pipe. Here, we ignore the extra complication of the effect of the rotation of one of the annuli on the fluid flow.
## 2 Solution
For an incompressible fluid, the velocity distribution obeys the continuity equation
$$๐ฏ=0,$$
(1)
in which case the Navier-Stokes equation for steady, viscous flow is
$$\rho (๐ฏ)๐ฏ=P+\eta ^2๐ฏ.$$
(2)
There are only three examples in which analytic solutions to this equation have been obtained when the nonlinear term $`(๐ฏ)๐ฏ`$ is nonvanishing .
We first review the simpler case of two-dimensional flow between parallel plates in sec. 2.1, and then take up the case of radial flow in sec. 2.2. We will find an analytic solution to the nonlinear Navier-Stokes equation (2) for radial flow, but this solution cannot satisfy the the boundary conditions,
$$๐ฏ(z=0)=0=๐ฏ(z=h),$$
(3)
that the flow velocity vanish next to the plates. However, in the linear approximation to eq. (2) we obtain an analytic form for the radial flow between two annular plates.
### 2.1 Two-Dimensional Flow between Parallel Plates
For guidance, we recall that an analytic solution is readily obtained for the related problem of two-dimensional viscous flow between two parallel plates. For example, suppose that the plates are at the planes $`z=0`$ and $`z=h`$, and that the flow is in the $`x`$ direction, i.e., $`๐ฏ=v(x,z)\widehat{๐ฑ}`$. The equation of continuity (1) then tells us that $`v/x=0`$, so that the velocity is a function of $`z`$ only,
$$๐ฏ=v(z)\widehat{๐ฑ}.$$
(4)
The $`z`$ component of the Navier-Stokes equation (2) reduces to $`P/z=0`$, so that the pressure is a function of $`x`$ only. The $`x`$ component of eq. (2) is
$$\frac{P(x)}{x}=\eta \frac{^2v(z)}{z^2}.$$
(5)
Since the lefthand side is a function of $`x`$, and the righthand side is a function of $`z`$, equation (5) can be satisfied only if both sides are constant. Supposing that the pressure decreases with increasing $`x`$, we write
$$\frac{P}{x}=\text{constant}=\frac{\mathrm{\Delta }P}{\mathrm{\Delta }x}>0.$$
(6)
Using the boundary conditions (3), we quickly find that
$$v(z)=\frac{\mathrm{\Delta }P}{\mathrm{\Delta }x}\frac{z(hz)}{2\eta }=6\overline{v}\frac{z}{h}\left(1\frac{z}{h}\right),$$
(7)
where the average velocity $`\overline{v}`$ is given by
$$\overline{v}=\frac{1}{h}_0^hv(z)๐z=\frac{\mathrm{\Delta }P}{\mathrm{\Delta }x}\frac{h^2}{12\eta }.$$
(8)
### 2.2 Radial Flow between Parallel Annular Plates
Returning to the problem of radial flow between two annular plates, we seek a solution in which the velocity is purely radial, $`๐ฏ=v(r,z)\widehat{๐ซ}`$. The continuity equation (1) for this hypothesis tells us that
$$\frac{1}{r}\frac{(rv)}{r}=0,$$
(9)
so that
$$๐ฏ=\frac{f(z)}{r}\widehat{๐ซ}.$$
(10)
Following the example of two-dimensional flow between parallel plates, we expect a parabolic profile in $`z`$ as in eq. (7),
$$f(z)z(hz),$$
(11)
which satisfies the boundary conditions (3).
Using the trial solution (10), the $`z`$ component of the Navier-Stokes equation (2) again tells us that the pressure must be independent of $`z`$: $`P=P(r)`$. The radial component of eq. (2) yields the nonlinear form
$$\eta r^2\frac{d^2f}{dz^2}+\rho f^2=r^3\frac{dP}{dr}.$$
(12)
The hoped-for separation of this equation can only be achieved if $`f(z)=F`$ is constant, which requires the pressure profile to be $`P(r)=A\rho F^2/2r^2`$. The boundary conditions (3) cannot be satisfied by this solution. Further, this solution exists only for the case that the pressure is increasing with increasing radius. The fluid flow must be then be inward, so the constant $`F`$ must be negative. The Navier-Stokes equation is not time-reversal invariant due to the dissipation of energy associated with the viscosity, and so reversing the velocity of a solution does not, in general, lead to another solution.
While we have obtained an analytic solution to the nonlinear Navier-Stokes equation (2), it is not a solution to the problem of radial flow between two annuli. It is hard to imagine a physical problem involving steady, radially inward flow of a long tube of fluid, to which the solution could apply.
Instead of an exact solution, we are led to seek an approximate solution in which the nonlinear term $`f^2`$ of eq. (12) can be ignored. In this case, the differential equation takes the separable form
$$\frac{d^2f}{dz^2}=\frac{r}{\eta }\frac{dP}{dr}=\text{constant}.$$
(13)
Following eq. (7) we write the solution for $`f`$ that satisfies the boundary conditions (3) as
$$f(z)=6\overline{f}\frac{z}{h}\left(1\frac{z}{h}\right),$$
(14)
where $`\overline{f}`$ is the average of $`f(z)`$ over the interval $`0zh`$. The part of eq. (13) that describes the pressure leads to the solution
$$P(r)=\frac{P_1\mathrm{ln}r_2/r+P_2\mathrm{ln}r/r_1}{\mathrm{ln}r_2/r_1},$$
(15)
where $`P_i=P(r_i)`$. Plugging the solutions (14) and (15) back into eq. (13), we find that
$$\overline{f}=\frac{h^2\mathrm{\Delta }P}{12\eta \mathrm{ln}r_2/r_1},$$
(16)
where $`\mathrm{\Delta }P=P_1P_2`$. Hence, the flow velocity is
$$๐ฏ(r,z)=\frac{z(hz)\mathrm{\Delta }P}{2\eta r\mathrm{ln}r_2/r_1}\widehat{๐ซ},$$
(17)
whose average with respect to $`z`$ is $`\overline{v}(r)=\overline{f}/r`$. As with all solutions to the linearized Navier-Stokes equation, the velocity is independent of the density.
For the approximate solution (17) to be valid, the term $`f^2\overline{f}^2`$ must be small in eq. (12), which requires
$$\frac{\rho h^4\mathrm{\Delta }P}{144\eta ^2r_1^2\mathrm{ln}r_2/r_1}1.$$
(18)
When this condition is not satisfied, the solution must include velocity components in the $`z`$ direction that are significant near the inner and outer radii, while the flow pattern at intermediate radii could be reasonably well described by eq. (17).
If one of the annuli is rotating at angular velocity $`\omega `$, the radial flow velocity should still be given approximately by eq. (17) so long as ฯr2
<
vยฏ(r2)=fยฏ/r2
<
๐subscript๐2ยฏ๐ฃsubscript๐2ยฏ๐subscript๐2\omega r_{2}\mathrel{\vbox{\kern 0.0pt\hbox{$<$} \kern 0.0pt\hbox{$\sim$} }}\bar{v}(r_{2})=\bar{f}/r_{2}.
|
warning/0006/astro-ph0006175.html
|
ar5iv
|
text
|
# Re-processing the Hipparcos Transit Data and Intermediate Astrometric Data of spectroscopic binaries. I. Ba, CH and Tc-poor S starsBased on observations from the Hipparcos astrometric satellite operated by the European Space Agency (ESA 1997)
## 1 Introduction
The chemical anomalies observed in several classes of late-type stars are due to mass transfer across a binary system. This scenario, first suggested by McClure (1983), holds for dwarf and giant barium stars (Pop.I G and K stars with overabundance of carbon and heavy elements like Sr and Ba produced by the s-process of nucleosynthesis), CH stars (the Pop.II analogs of Ba stars), and extrinsic S stars (late-type giants characterized by ZrO bands and no Tc lines, an element with no stable isotopes). The set of spectroscopic orbits available for those chemically-peculiar red stars (CPRS) has been considerably enlarged thanks to a decade-long radial-velocity monitoring. Such orbits are now available for 63 giant barium stars and 19 extrinsic S stars (Jorissen et al. (1998)), 21 dwarf barium and subgiant CH stars (McClure (1997);North, priv. comm.) and 8 CH stars (McClure & Woodsworth (1990)). Most of them were not available to the Hipparcos reduction consortia. The astrometric parameters (parallaxes and proper motions) published in the Hipparcos catalogue (ESA (1997)) were therefore derived from a single star solution (the so-called โ5-parameter modelโ). Since many of these CPRS have orbital periods in the range 1 โ 5 yr, their unaccounted orbital motion may seriously confuse the parallax and proper motion as determined by the Hipparcos consortia. Thus, unrecognized orbital motions might introduce a so-called โcosmic errorโ (of a systematic nature) in the proper motion (Wielen (1997); Wielen et al. (1999)). Parallaxes may also possibly be in error, especially for binaries with orbital periods close to 1 yr.
The conclusions of previous studies inferring the kinematics and absolute magnitudes of CPRS (Bergeat & Knapik (1997); Mennessier et al. (1997); Van Eck et al. (1998)) from the data of the Hipparcos catalogue may possibly be affected by these systematic errors. It is therefore of importance to re-evaluate the astrometric parameters of CPRS with a model correctly accounting for the orbital motion (the so-called โ12-parameter modelโ, reducing to 9 parameters if the spectroscopic orbital parameters $`P,T`$ and $`e`$ are fixed), and to compare these values with those from the Hipparcos catalogue to identify the stars most affected by these systematic errors.
This comparison should allow to assess the reliability of the kinematic properties and absolute magnitudes of CPRS formerly derived from the Hipparcos-catalogue values. As a by-product, the astrometric orbit has been obtained in many cases, leading to an estimate of the masses through the ratio $`M_2/[1+(M_1/M_2)]^2f(M)/\mathrm{sin}^3i`$ (derived by combining the orbital inclination from the astrometric orbit with the mass function $`f(M)`$ from the spectroscopic orbit when only one spectrum is observable).
From the operational point of view, the numerical re-processing of the raw Hipparcos data (Intermediate Astrometric Data or Transit Data; see Sect. 2.1) presented in this paper is innovative on the following points: (i) a global optimization technique is used to minimize the objective function in the 12-parameter (or, for spectroscopic binaries, 9-parameter) space, and (ii) a change of independent variable allows to derive always positive parallaxes, contrarily to the situation prevailing in the Hipparcos catalogue. The tool developed here for CPRS may in the future be applied to any other binary star to re-process the Hipparcos data when new orbits appear. The second paper of this series will be devoted to the re-processing of spectroscopic binaries involving a giant component (Boffin et al. (1993)). Another is devoted to Procyon (Girard et al. (2000)). We encourage interested readers to communicate to the authors other spectroscopic binaries whose Hipparcos data would need to be reprocessed.
## 2 Numerical methods
In order to allow some re-processing of the Hipparcos data, the Hipparcos consortia provided the users with two kinds of data: the Transit Data (TD) and the Intermediate Astrometric Data (IAD). TD (Quist & Lindegren (1999)) are a by-product of the analysis performed by the NDAC consortium, and merge astrometric and photometric data. They basically provide the signal as modulated by the grid in front of the detector. They are therefore especially well suited to the analysis of visual double or multiple systems, since the separation, position angle and magnitudes of the various components may in principle be extracted from the TD. For that reason, TD are not provided for all Hipparcos entries, but only for those stars known or suspected of being double or multiple systems. IAD are lower level data which provide astrometric information only (the starโs abscissa along a reference great circle, and the pole of that circle on the sky; van Leeuwen & Evans (1998)), but for every Hipparcos entry. In the following, we describe how IAD and TD (when available) have been used to derive the astrometric parameters of the CPRS that have a known spectroscopic orbit. When TD are available, the astrometric parameters may be derived independently from the two data sets, and allow an interesting internal consistency check. In both cases, the method proceeds along the following steps:
* definition of an objective function to minimize
* computation of the objective function from the astrometric parameters
* global minimization of the objective function with the simulated annealing algorithm
* local minimization of the objective function (with a quasi-Newton method)
* comparison of the values of the objective function for the Hipparcos 5-parameter model and for our 9-parameter model, and application of an F-test to evaluate whether the 9-parameter model is significantly better than the Hipparcos solution
* check of the astrophysical consistency of the 9-parameter solution from some a priori knowledge of the properties of the system (i.e., component masses)
These steps are discussed in turn in the remaining of this section, first for IAD, and then for TD.
### 2.1 Intermediate Astrometric Data
#### 2.1.1 Objective function
The first step in any data-fitting problem is to set up an objective function in order to compare different solutions (corresponding to different values of the model parameters). This function is usually constructed in such a way that its lowest value corresponds to the best solution. The problem then reduces to minimizing that function. The IAD already include some corrections (e.g. aberration and satellite-attitude corrections) and they are no longer stricto sensu raw data. They provide the abscissa residuals along a reference great circle for the Hipparcos 5-parameter solution (i.e. $`\alpha _0,\delta _0,\varpi ,\mu _\alpha ,\mu _\delta `$, respectively: the position in right ascension and declination in the equinox 2000.0 at the epoch 1991.25, the parallax, the proper motions in right ascension and declination; following the practice of the Hipparcos catalogue, $`\alpha _0^{}=\alpha _0\mathrm{cos}\delta `$ and $`\mu _\alpha ^{}=\mu _\alpha \mathrm{cos}\delta `$ will be used instead of $`\alpha _0`$ and $`\mu _\alpha `$). By slightly modifying these 5 parameters and by adding 7 orbital parameters (Binnendijk (1960)), the aim is to further reduce the abscissa residuals $`\mathrm{\Delta }v`$ below the values obtained from the Hipparcos 5-parameter model. In the case of the Hipparcos Intermediate Astrometric Data, the objective function is the $`\chi ^2`$ expressed by Eq. 17.12 of Vol.3 in ESA (1997):
$$\chi ^2=(\mathrm{\Delta }v\underset{k=1}{\overset{M}{}}\frac{v}{p_k}\mathrm{\Delta }p_k)^\text{t}V^1(\mathrm{\Delta }v\underset{k=1}{\overset{M}{}}\frac{v}{p_k}\mathrm{\Delta }p_k),$$
(1)
where the superscript $`t`$ means transposed, $`\mathrm{\Delta }v`$ are the abscissa residuals provided by the IAD file and corresponding to the Hipparcos 5-parameter solution, and $`\frac{v}{p_k}`$ is the partial derivative of the abscissa with respect to the $`k`$-th parameter expressing how the abscissa residual varies when a correction $`\mathrm{\Delta }p_k`$ is applied to the value of the $`k`$-th parameter with respect to the Hipparcos solution. $`M`$ is the number of parameters retained in the solution, and $`V^1`$ is the inverse of the covariance matrix of the observations given by:
$$V=\left(\begin{array}{cccc}V_1& 0& \mathrm{}& 0\\ 0& V_2& \mathrm{}& \mathrm{}\\ \mathrm{}& \mathrm{}& \mathrm{}& 0\\ 0& \mathrm{}& 0& V_n\end{array}\right)$$
with
$$V_j=\left(\begin{array}{cc}\sigma _{F_j}^2& \rho \sigma _{F_j}\sigma _{N_j}\\ \rho \sigma _{F_j}\sigma _{N_j}& \sigma _{N_j}^2\end{array}\right)$$
(2)
If an observation $`j`$ was processed by the two consortia (FAST and NDAC), the residuals obtained by the two reduction consortia are correlated and $`V_j`$ is the $`2\times 2`$ variance-covariance matrix for measurement $`j`$. On the other hand, when only one consortium processed observation $`j`$, $`V_j`$ reduces to one number<sup>1</sup><sup>1</sup>1In order to have a unique expression for $`\chi ^2`$ regardless of the dataset available (FAST, NDAC or both), we always write $`V_j`$ as a 2 by 2 matrix. When, for whatever reason, a measurement from only one consortium is used, that observation is duplicated and its weight divided by two ($`V_j=\text{diag}(2\sigma _j^2,2\sigma _j^2)`$). There is no change on the actual value of $`\chi ^2`$ but its expression is easier (though not faster!) to evaluate. (the estimated uncertainty $`\sigma _j`$ of the measurement). $`\mathrm{\Delta }v`$, together with $`\frac{v}{p_k}`$ ($`k=1`$ to 5, with $`p_1\alpha _0^{},p_2\delta ,p_3\varpi ,p_4\mu _\alpha ^{},p_5\mu _\delta `$), the original astrometric parameters as well as $`\rho ,\sigma _{F_j}`$ and $`\sigma _{N_j}`$ are given in the IAD file. In order to evaluate $`\chi ^2`$ (Eq. 1) for an orbital model, expressions for the partial derivatives of $`v`$ with respect to the orbital parameters are required. They can be expressed as a function of the partial derivatives of $`v`$ with respect to $`\alpha _0^{}`$ and $`\delta _0`$ as follows (Hipparcos catalogue, Vol. 3, Eq. $`17.15`$):
$$\frac{v}{o}=\frac{v}{\alpha _0^{}}\frac{\xi }{o}+\frac{v}{\delta _0}\frac{\eta }{o}$$
where $`o`$ is any orbital parameter and
$$\begin{array}{c}\xi =\alpha _0^{}+\mu _\alpha ^{}(tt_0)+RP_\alpha \varpi +y,\hfill \\ \eta =\delta _0+\mu _\delta (tt_0)+RP_\delta \varpi +x.\hfill \end{array}$$
(3)
In the above expression, $`\xi `$ and $`\eta `$ represent the Cartesian coordinates of the observed component on the plane tangent to the sky at the position ($`\alpha _0,\delta _0`$). They combine the displacements due to the proper motion, the orbital motion and the parallax. $`R`$ is the radius vector of the Earthโs orbit in A.U. at time $`t`$. $`P_\alpha `$ and $`P_\delta `$, the parallax factors, are given by Binnendijk (1960):
$`P_\alpha `$ $`=`$ $`\mathrm{cos}ฯต\mathrm{cos}\alpha \mathrm{sin}\mathrm{}\mathrm{sin}\alpha \mathrm{sin}\mathrm{}`$
$`P_\delta `$ $`=`$ $`(\mathrm{sin}ฯต\mathrm{cos}\delta \mathrm{cos}ฯต\mathrm{sin}\alpha \mathrm{sin}\delta )\mathrm{sin}\mathrm{}`$
$`\mathrm{cos}\alpha \mathrm{sin}\delta \mathrm{cos}\mathrm{}`$
where $`\mathrm{}`$ and $`ฯต`$ are respectively the longitude of the Sun and the obliquity of the ecliptic, both at time $`t`$.The variables $`x`$ and $`y`$ describe the apparent orbit (i.e., projected on the plane orthogonal to the line of sight) of the observable component around the center of mass of the system. They are usually expressed in terms of the Thiele-Innes constants $`A,B,F,G`$ of the photocentric orbit as (Heintz (1978))
$$\begin{array}{c}x=AX+FY\hfill \\ y=BX+GY\hfill \end{array}$$
(4)
with
$`X`$ $`=`$ $`\mathrm{cos}Ee`$
$`Y`$ $`=`$ $`\sqrt{1e^2}\mathrm{sin}E,`$
where $`E`$ is the eccentric anomaly and $`(X,Y)`$ are the coordinates in the true orbit. It should be noted that $`(x,y)`$ and $`(X,Y)`$ are referred to a Cartesian system with $`x`$ pointing towards the North (Heintz (1978)), contrary to $`(\xi ,\eta )`$ where $`\xi `$ points towards increasing right ascensions. For all the systems considered in the present paper, the magnitude difference between the two components is larger than the Hipparcos detection threshold (since the companion to CPRS is a cool white dwarf; Sect. 5 and Jorissen et al. (1998)). Thus, one can safely assume that there is no light coming from the secondary. Hence, the orbit of the photocenter of the primary component is the same as the absolute orbit of the primary around the center of mass of the system.
At this point, it is important to realize that the Hipparcos 5-parameter solution, that is used as a starting point for the new 12-parameter solution, is in fact equivalent to a 12-parameter solution where the semi-major axis of the orbit ($`a_0`$) is null. It is therefore enough to consider in Eq. 1 the correction term relative to the semi-major axis of the orbit (and in fact $`\mathrm{\Delta }a_0=a_0`$, since the initial value of $`a_0`$ is null). The other orbital parameters $`i,\omega ,\mathrm{\Omega },e,P`$ and $`T`$ enter Eq. 1 only through the partial derivatives $`\frac{\xi }{a_0}`$ and $`\frac{\eta }{a_0}`$ \[equal, respectively, to $`\frac{y}{a_0}`$ and $`\frac{x}{a_0}`$ according to Eq. 3\] entering $`\frac{v}{a_0}`$. These other orbital parameters do not require explicit correction terms in Eq. 1 (and their starting values would be ill-defined anyway). Hence, the orbital solution is the one minimizing
$$\chi ^2=\mathrm{\Xi }^\text{t}V^1\mathrm{\Xi }$$
(5)
where
$`\mathrm{\Xi }`$ $`=`$ $`\mathrm{\Delta }v{\displaystyle \underset{k=1}{\overset{5}{}}}{\displaystyle \frac{v}{p_k}}\mathrm{\Delta }p_k({\displaystyle \frac{v}{p_1}}{\displaystyle \frac{\xi }{a_0}}+{\displaystyle \frac{v}{p_2}}{\displaystyle \frac{\eta }{a_0}})a_0`$ (6)
is a vector of dimension N (N is the number of observations), and $`\xi `$ and $`\eta `$ are functions of $`a_0,i,\omega ,\mathrm{\Omega },e,P`$ and $`T`$, and $`\frac{\xi }{a_0}\frac{y}{a_0},\frac{\eta }{a_0}\frac{x}{a_0}`$. Thus, the 12 parameters enter the evaluation of $`\chi ^2`$ although there are only six correction terms subtracted from $`\mathrm{\Delta }v`$.
Our experience has shown that, except for very special cases (i.e., parallaxes and semi-major axes larger than about 20 mas, orbital periods significantly different from 1 year but smaller than 3 years; one example is HIP 50805 in Table 1), astrometric orbits could not be derived from the IAD without an a priori knowledge of some of the orbital elements, for instance the spectroscopic ones ($`e`$, $`P`$ and $`T`$). However, it appeared that fixing $`\omega `$ at the value derived from the spectroscopic orbit often led to orbital inclinations $`i`$ unrealistically close to zero for spectroscopic binaries with radial velocity variations. Leaving the parameter $`\omega `$ free removes this difficulty, and offers moreover a way to check the consistency of the astrometric solution, since the astrometric $`\omega `$ should be consistent with its spectroscopic value.
As far as outliers are concerned, we almost always keep the same data set as that used by FAST and/or NDAC (Vol. 3, Sect. 17.6;ESA (1997)). IAD that were not considered by FAST or NDAC (i.e., with the IA2 field set to โfโ or โnโ) were thus not included in our re-processing either. In a few cases, we noticed that because of the orbital contribution, some observations yield residuals larger than $`3\sigma `$ of the residuals. In these cases, these observations were removed and the fit re-iterated. We never had to iterate more than twice to remove all outliers.
#### 2.1.2 Forcing parallaxes to be positive
One of the five astrometric parameters entering $`\chi ^2`$ in Eq. 1 is the parallax $`\varpi `$. With no other prescriptions as those described in Sect. 2.1.1, the minimization process may very well end up with a negative parallax. Indeed, negative parallaxes are not rare in the Hipparcos and Tycho catalogues.
Parallaxes cannot only be seen as inverse distances (which are defined positive) but also as the semi-major axis of the parallactic ellipse (see Eq. 3). The direction of motion along the parallactic ellipse is of course imposed by the annual revolution of the Earth around the Sun, regardless of the actual dimension of the parallactic ellipse or of the observational uncertainties. In that sense, the parallactic ellipse is oriented, and negative parallaxes can be seen as corresponding to a parallactic ellipse covered in the wrong direction. That constraint being of physical nature, one should seek to fulfill it. In this section, we present a method which always delivers positive parallaxes. This method is especially useful for stars like those Mira variables or carbon stars that came out with large negative parallaxes in the Hipparcos Catalogue. In those cases, forcing the parallax to be positive has a strong impact on the derived proper motion (Pourbaix et al., in preparation), which may be supposed to be better determined with a physically- sound model yielding positive parallaxes. However, the major drawback of the method is that the errors on the parallax do no longer follow a normal distribution. Therefore, the use of the parallaxes provided by this paper for e.g., luminosity calibrations should be done with care to avoid biases.
In order to force the parallax to be positive โ and at the same time avoiding the difficulties inherent to any constrained minimization techniques โ one may replace the constrained variable (the parallax $`\varpi `$) by an unconstrained one<sup>2</sup><sup>2</sup>2We are indebted to A. Albert for suggesting this trick. An appropriate choice appears to be
$$\varpi ^{}=\mathrm{log}\varpi $$
(7)
which is $`C^{\mathrm{}}`$ between $`]0,+\mathrm{}[`$ and $``$. The variable $`\varpi ^{}`$ is in fact equivalent to the distance modulus $`mM`$ since
$$mM=55\varpi ^{}.$$
(8)
The variable $`\varpi ^{}`$ is thus used in the minimization process instead of $`\varpi `$ as one of the $`p_k`$ parameters entering Eq. 1 \[with $`\frac{v}{\varpi ^{}}=\frac{v}{\varpi }\varpi \mathrm{ln}10`$\]. At the end of the minimization, Eq. 7 is reversed and the parallax $`\varpi `$ is derived from $`\varpi ^{}`$. For any real number $`\varpi ^{}`$, $`\varpi `$ necessarily lies in between 0 and $`\mathrm{}`$. Such a substitution is legitimate, since the parallax is not a directly measured quantity, but rather one among many parameters used in a model fit to the observations. Moreover, since maximum-likelihood estimators<sup>3</sup><sup>3</sup>3$`\chi ^2`$ is a maximum-likelihood estimator provided that the measurement errors are normally distributed enjoy the invariance property (see e.g., Mood et al. (1974), p. 284), $`\varpi `$\- and $`\varpi ^{}`$-fitting must yield identical results for those cases where $`\varpi `$-fitting yielded a positive parallax (since the $`\varpi ^{}=\mathrm{log}\varpi `$ transform may be used when $`\varpi >0`$ to extract $`\varpi ^{}`$).
Most of the objects considered in the present paper have large parallaxes that would have come out positive by a direct fit of $`\varpi `$ anyway. Thus, in the present case, the fitting of $`\varpi ^{}`$ (instead of $`\varpi `$) does not represent so much of an improvement. Nevertheless, the procedure of $`\varpi ^{}`$-fitting has been introduced here for the sake of generality.
The price to pay is, however, that the errors on $`\varpi `$ do not any longer follow a normal distribution. Moreover, the confidence interval of the parallax is no more symmetric. It has been estimated by the following expression:
$$\begin{array}{c}\widehat{\varpi }^{}\sigma \varpi ^{}\widehat{\varpi }^{}+\sigma \hfill \\ 10^{\widehat{\varpi }^{}\sigma }\varpi 10^{\widehat{\varpi }^{}+\sigma }\hfill \end{array}$$
(9)
where $`\widehat{\varpi }^{}`$ is the value resulting from the $`\chi ^2`$\- minimization, and $`\sigma `$ its estimated uncertainty. It should be noted that the errors on $`\varpi ^{}`$ do not follow a normal distribution either, since the normality is only guaranteed for โlinear modelsโ (Mood et al. (1974)) and $`\mathrm{\Xi }`$ (Eq. 5) does not depend linearly upon $`\varpi ^{}`$. Therefore, the confidence interval corresponding to a given probability level is in general not symmetric around $`\varpi ^{}`$. However, to ease the computations, the uncertainty $`\sigma `$ on $`\varpi ^{}`$ will be computed from the inverse of the Fisher information matrix at the point minimizing $`\chi ^2`$ (see Sect. 2.1.3). This procedure implicitly assumes that the model is linear in the vicinity of the minimum, so that the adopted confidence interval is in fact one that is symmetric around $`\varpi ^{}`$.
The second expression in Eq. 9 clearly shows that the parallax is positive everywhere in the confidence interval, which would not be guaranteed in a constrained minimization of $`\chi ^2`$. That important property illustrates the superiority of this approach with respect to the constrained minimization.
#### 2.1.3 $`\chi ^2`$ minimization
If the $`\chi ^2`$ expressed by Eq. 5 were a quadratic expression of the unknown parameters $`p_k`$, its unique minimum could be found from the solution of a set of linear equations. However, the parameters $`i`$, $`\omega `$ and $`\mathrm{\Omega }`$ enter $`\chi ^2`$ in a highly non-linear way, so that the function expressing $`\chi ^2`$ in the 9-parameter space may have several local minima, and finding its global minimum is a much more arduous task.
Faced with such situations, one of us (DP) has already successfully worked out global optimization techniques such as simulated annealing (Pourbaix (1994, 1998b)). Practical details about the implementation of the method to minimize the objective function $`\chi ^2`$ \[Eq. 5\] in the working space $`^9`$ may be found in Pourbaix (1998a). Simulated annealing being a heuristic method, one can only prove its convergence to the global minimum after an infinite time (which we cannot afford). We thus stop the procedure after a finite time. In order to nevertheless have a good chance to obtain (a neighborhood of) the global minimum, we repeat 40 times this highly non deterministic minimization process. The best solution ever met (i.e., the one leading to the smallest $`\chi ^2`$ value) is finally adopted. Once (a neighborhood of) the global minimum is thus obtained, it is tuned with the BFGS quasi-Newton algorithm (Dennis Jr. & Schnabel (1995)).
Unlike the Levenberg-Marquardt (Marquardt (1963)) minimization algorithm, BFGS does not return the covariance matrix of the model parameters. The inverse of the Fisher information matrix at the minimum is therefore used as the best estimate of that covariance matrix (Pourbaix (1994)).
The whole procedure has been applied separately on the data from the FAST consortium only, from NDAC only and from both combined, thus resulting in three different solutions, hopefully consistent with each other.
In a few instances, the solution obtained from the combined FAST+NDAC data set turns out to be very close to either the FAST or NDAC solution, but FAST and NDAC taken separately yield rather different solutions. That situation probably reflects the very different weights attributed to the two data sets for that particular object in the merging process applied to produce the output catalogue. For our analysis we always keep the covariance matrices of the observations as they are given in the electronic version of the catalogue.
As pointed out by an anonymous referee, in the case where all the Campbell elements ($`a_0`$, $`i`$, $`\omega `$ and $`\mathrm{\Omega }`$) are extracted from a fit to the astrometric data, they can advantageously be replaced by the Thiele-Innes elements ($`A`$, $`B`$, $`F`$ and $`G`$) so that $`\chi ^2`$ becomes a quadratic function of the model parameters. The minimum of $`\chi ^2`$ can then be found analytically and no minimization (neither global nor local) technique is needed. For the sake of generality, we nevertheless use the Campbell set (and thus the minimization scheme) because this more general scheme allows, if necessary, to easily incorporate external constraints (like for instance the knowledge of $`i`$ for eclipsing binaries, or $`\omega `$ from the spectroscopic orbit; see, however, the comment about fixing $`\omega `$ after Eq. 6 in Sect. 2.1.1). Such additional constraints would be much more difficult to impose through the Thiele-Innes elements.
#### 2.1.4 F-test: 5-parameter vs orbital model
The introduction of more free parameters in the orbital model as compared to the single star solution necessarily leads to a reduction of the objective function. To evaluate whether this reduction is statistically significant โ or, equivalently, whether the orbital solution represents a significant improvement over the 5-parameter Hipparcos solution โ requires the use of an F-test.
The method used here is inspired from the test devised by Lucy & Sweeney (1971). If $`\chi _{\mathrm{Hip}}^2`$ and $`\chi _{\mathrm{orb}}^2`$ denote the residuals for the Hipparcos 5-parameter model and for the orbital model (with 9 free parameters), respectively, the efficiency of the additional 4 parameters in reducing $`\chi _{\mathrm{orb}}^2`$ below $`\chi _{\mathrm{Hip}}^2`$ may be measured by the ratio:
$$F=\frac{N9}{4}\frac{\chi _{\mathrm{Hip}}^2\chi _{\mathrm{orb}}^2}{\chi _{\mathrm{orb}}^2},$$
(10)
where $`N`$ is the number of available measurements.
If the hypothesis that there is no orbital motion (i.e., $`a_0=0`$ in Eq. 5) is correct, then it may be shown (Bevington & Robinson (1992)) that $`F`$ follows a Snedecor $`F_{\nu _1,\nu _2}`$ distribution with $`\nu _1=4`$ and $`\nu _2=N9`$ degrees of freedom. Thus, if Eq. 10 yields $`F=\widehat{F}`$, then, on the assumption that there is no orbital motion, the probability that $`F`$ could have exceeded $`\widehat{F}`$ is
$$\alpha =\text{Prob}(F>\widehat{F})Q(\widehat{F}|\nu _1,\nu _2).$$
(11)
In other words, $`\alpha `$ is the first risk error of rejecting the null hypothesis that the orbital and 5-parameter models are identical while it is actually true.
The residuals given in the IAD files always relate to a 5-parameter solution, even when a more sophisticated model (the so-called โaccelerationโ 7- or 9-parameter models, or even orbital model) was published in the Hipparcos catalogue (see Table 1). For those cases, the $`\alpha `$ value listed in Table 1 is always close to zero, although it does not really characterize the improvement of the orbital solution with respect to the solution retained in the Hipparcos catalogue (which goes already beyond a 5-parameter model).
#### 2.1.5 Astrophysical consistency of the orbital solution
The minimization process will yield a solution in all cases, but that solution may not be astrophysically relevant. A statistical check of the significance of the orbital solution, based on the $`F`$-test, has been presented in Sect. 2.1.4. In this section, two criteria testing the validity of the orbital solution on astrophysical grounds are presented.
The first test is based upon the identity
$$\frac{a_0\mathrm{sin}i}{\varpi }=\frac{PK_1\sqrt{1e^2}}{2\pi },$$
(12)
where $`K_1`$ is the radial-velocity semi-amplitude of the visible component. The left-hand side of the above identity entirely depends upon astrometric parameters, whereas its right-hand side contains only parameters derived spectroscopically.
This test has the advantage of being totally independent of any assumptions. However, it involves the orbital inclination which is not always very accurately determined (see Sect. 3.2), so that the above identity may not always be very constraining considering the often large uncertainty on $`i`$.
A somewhat more constraining identity to assess the astrophysical plausibility of the computed astrometric orbit is the following:
$$\frac{a_0}{\varpi }=P^{2/3}\frac{M_2}{(M_1+M_2)^{2/3}},$$
(13)
where $`M_1`$ and $`M_2`$ are the masses of the visible and invisible components, respectively. However, it relies upon assumptions regarding the stellar masses. In some cases (e.g., dwarf barium stars), the mass of the observable component may be estimated directly from the spectroscopically-derived gravity and from an estimate of the stellar radius from the spectral type (North et al. (1999)). For the other samples, an average mass derived from a statistical analysis of the spectroscopic mass function (Jorissen et al. (1998)) is adopted. For all the samples considered here, the unseen component is almost certainly a white dwarf (the only possible exception being the Tc-poor S star HIP 99312 = HD 191589), whose mass may be taken as $`0.62\pm 0.04`$ M (Jorissen et al. (1998)).
Because of the assumptions involved, this test is only used as a guide to identify astrophysically-unplausible solutions. It turns out that such cases are generally those with large error bars or with inconsistent N, F and A solutions, thus providing further arguments not to retain those solutions. In very few cases (HIP 36042, 53763 and 60299), valid data yielded solutions not consistent with Eq. 13. Those cases were nevertheless kept in our final list.
### 2.2 Transit Data
Unlike the IAD, TD are only available for a small subset of the Hipparcos catalogue, e.g., for those stars that were known to be (or suspected of being) double or multiple systems at the time of the data reduction by the Hipparcos consortia. TD are a by-product of (or, more precisely, an input for) the multiple-star processing by the NDAC consortium. Another difference with respect to the IAD concerns photometry. Whereas IAD contain astrometric information only, the brightness of the (different) observable component(s) of the system can be retrieved from the TD.
In the most general case (Quist & Lindegren (1999)), each entry in the TD file corresponds to five numbers $`b_1`$, โฆ, $`b_5`$ which represent the coefficients of the first terms in the Fourier series modeling the observed signal as modulated by the detector grid:
$`b_1`$ $`=`$ $`{\displaystyle \underset{i}{}}I_i`$
$`b_2`$ $`=`$ $`\stackrel{~}{M}{\displaystyle \underset{i}{}}I_i\mathrm{cos}\varphi _i`$
$`b_3`$ $`=`$ $`\stackrel{~}{M}{\displaystyle \underset{i}{}}I_i\mathrm{sin}\varphi _i`$
$`b_4`$ $`=`$ $`\stackrel{~}{N}{\displaystyle \underset{i}{}}I_i\mathrm{cos}(2\varphi _i)`$
$`b_5`$ $`=`$ $`\stackrel{~}{N}{\displaystyle \underset{i}{}}I_i\mathrm{sin}(2\varphi _i)`$
where $`I_i`$ is the intensity of the $`i`$-th component of the system, and the phase $`\varphi _i`$ corresponds to its abscissa along the reference great circle. $`\stackrel{~}{M}=0.7100`$ and $`\stackrel{~}{N}=0.2485`$ are the adopted reference values for the modulation coefficients of the first and second harmonics (Vol. 1, Sect. 2.9; ESA (1997)). In terms of the Cartesian coordinates $`(\xi _i,\eta _i)`$ (see Eq. 3) of component $`i`$ in the plane tangent to the sky at the reference point specified in the TD file, the phase $`\varphi _i`$ writes
$$\varphi _i=f_x\xi _i+f_y\eta _i+f_p(\varpi \varpi _{\mathrm{ref}}).$$
(14)
The reference point is assigned an arbitrary parallax $`\varpi _{\mathrm{ref}}`$ and proper motion, as given in the TD file. The phase derivatives $`f_x,f_y`$ and $`f_p`$ with respect to $`\xi ,\eta `$ and $`\varpi `$ are also provided by the TD file.
For the SB1 systems we are interested in, the situation simplifies a lot since $`I_2`$ may be taken equal to 0. The above system of equations is rank-deficient. From the second and third equations, one can rewrite:
$$\stackrel{o}{\varphi }=\mathrm{arg}(b_2,b_3)$$
(15)
and also express the uncertainty on $`\varphi `$ as
$$\sigma _\varphi ^2=\left(\frac{\stackrel{o}{\varphi }}{b_2}\right)^2\sigma _{b_2}^2+\left(\frac{\stackrel{o}{\varphi }}{b_3}\right)^2\sigma _{b_3}^2.$$
With the above definitions, the objective function whose minimum is sought is given by:
$$D=\frac{1}{N}\underset{k=1}{\overset{N}{}}\sigma _{\varphi _k}^2(\stackrel{o}{\varphi _k}\varphi _k)^2$$
(16)
where $`N`$ is the number of TD. In the above expression, $`\stackrel{o}{\varphi _k}`$ is the observed phase at time $`t_k`$ as derived from Eq. 15, whereas $`\varphi _k`$ is the phase computed from Eqs. 3 and 14 (noting that $`f_xRP_\alpha +f_yRP_\delta f_p`$) for a given set of astrometric and orbital parameters.
The remaining of the method follows the same steps as described in relation with the IAD, i.e. global and local optimization, positiveness of the parallax, โฆ
## 3 Results
Table 1 lists the various samples that have been considered, namely dwarf barium and subgiant CH stars (McClure (1997); North, priv. comm.), mild and strong barium stars (Jorissen et al. (1998)), Tc-poor S stars (Jorissen et al. (1998)) and CH stars (McClure & Woodsworth (1990)). Their spectroscopic orbital parameters have been taken from the reference quoted.
Table 1 also provides various parameters that allow either to assess the quality of the derived orbital solution or to identify the reason why a reliable orbital solution could not be derived for some of the spectroscopic binaries considered. The following parameters potentially control our ability to derive an astrometric orbit from the IAD, with favorable circumstances being mentioned between the parentheses: the parallax (column 3; large parallax), the number of available IAD measurements (FAST+NDAC; column 4; large number of measurements), the ecliptic latitude (column 5; this parameter may play a role since it controls how different the orientations of the reference great circles are; favorable cases have absolute values larger than 47), the orbital period (column 6; in the range 1 โ 3 yr to ensure a good sampling of the orbit), the orbital eccentricity (column 7; low eccentricity).
The following columns characterize the quality of the solution obtained from the minimization process. Column 11 provides $`\chi ^2/(N_{\mathrm{IAD}}9)`$, which should be of the order of unity if the internal error $`\sigma `$ (Eq. 2) on the abscissae has been correctly evaluated by the reduction consortia and if the model provides an adequate representation of the data. The first risk error associated with rejecting the null hypothesis that the orbital and Hipparcos 5-parameter solutions are identical (Eq. 11) is given in column 12. Low values of $`\alpha `$ are generally associated with Hipparcos solutions of the G, X or O types (as listed in column 9; see the caption to Table 1 for more details), since the orbital motion is then large enough to have been noticed already by NDAC or FAST. Columns 13 and 14 compare the $`a_0/\varpi `$ ratio derived from the orbital solution to its expected value from Eq. 13. In columns 11 to 13, the data refer to the orbital solution obtained by combining NDAC and FAST data. For dwarf barium stars, the masses used to estimate $`a_0/\varpi `$ according to Eq. 13 are $`M_2=0.62\pm 0.04`$ M, whereas $`M_1`$ is derived from the spectroscopic gravity, with an estimated error of 0.05M (North, priv. comm.). According to the statistical analysis of the spectroscopic mass functions performed by Jorissen et al. (1998), $`M_1`$ and $`M_2`$ pairs (expressed in M) of ($`1.7\pm 0.2,0.62\pm 0.04`$), ($`2.1\pm 0.2,0.62\pm 0.04`$) and ($`1.8\pm 0.2,0.62\pm 0.04`$) have been adopted for strong barium stars, mild barium stars and Tc-poor S stars, respectively. The same analysis performed by McClure & Woodsworth (1990) for CH stars yielded $`M_1=1.0\pm 0.1`$ M and $`M_2=0.62\pm 0.04`$ M.
Astrometric orbits were accepted when $`\alpha `$ is smaller than 10 per cent and the expected $`a_0/\varpi `$ value falls within the $`1\sigma `$ confidence interval. A few cases not fulfilling these criteria were nevertheless accepted after visual inspection of the orbital arc.
Examination of Table 1 reveals that the following criteria need to be fulfilled in order to be able to extract a reliable astrometric orbit from the Hipparcos data: $`\varpi 5`$ mas, $`P10`$ yr, $`N25`$. The success rate is as follows for the various samples: dwarf barium stars (9/13), mild barium stars (3/24), strong barium stars (6/26), Tc-poor S stars (3/10) and CH stars (2/8). The high success rate for dwarf barium stars naturally results from the fact that these dwarf stars are on average closer from the sun than the giant stars.
Table 2 lists the astrometric and orbital parameters for the reliable orbits according to the criteria discussed above. The results from the different processing modes are collected in Table 2, according to the symbol given in column 2: H/S refers to the parameters from the Hipparcos catalogue and from the spectroscopic orbit (on that line $`a_0/\varpi `$ is the semi-major axis in A.U. estimated from Eq. 13), F refers to the processing of the IAD from FAST only, N to the IAD from NDAC only, A from the processing of the combined FAST/NDAC data set, O to the orbital parameters from the DMSA/O, and T to the parameters resulting from the TD.
Most of the retained orbits are indeed characterized by $`\chi ^2/(N9)`$ values of the order of unity, as expected. The first-risk errors $`\alpha `$ are not always close to 0, but if the derived value for the semi-major axis is in good agreement with its expected value, that agreement has been considered as sufficient for retaining the orbit. The only cases where the reverse situation occurs (small $`\alpha `$ but discrepant $`a_0/\varpi `$) are the dwarf barium star HIP 60299, the mild barium star HIP 36042 and the CH star HIP 53763. Although the orbit of the latter is not well defined, it has been kept in our final list to illustrate the large uncertainty on $`\varpi `$ resulting from an orbital period close to 1 yr (Sect. 4).
The F, N and A solutions for the retained orbits are also generally in good agreement, the only exceptions being the dwarf Ba stars HIP 62409 and HIP 116233, and the mild Ba star HIP 117607. However, the model parameters of these systems are highly correlated, and the different measurement errors in the different data sets thus drive the solution in different directions. This statement may be expressed in a quantitative way using the concept of efficiency $`ฯต`$ introduced by Eichhorn (1989) and Pourbaix & Eichhorn (1999). It is defined as
$$ฯต=\sqrt[p]{\frac{_{k=1}^p\lambda _k}{_{k=1}^pq_{kk}}},$$
(17)
where $`\lambda _k`$ are the eigenvalues of the covariance matrix of the estimated parameters, $`q_{kk}`$ are its diagonal elements, and $`p`$ denotes the number of parameters in the model. If $`ฯต`$ is close to unity, there is obviously little correlation between the parameters. For the combined NDAC+FAST solution, it amounts to 0.32, 0.21 and 0.42 in the case of HIP 62409, HIP 116233 and HIP 117607, respectively, thus translating some degree of correlation between the model parameters.
In Table 2, the uncertainty on $`a_0/\varpi `$ has been computed by combining the upper and lower limits on $`a_0`$ and $`\varpi `$, thus neglecting any possible correlation between these two quantities (which is generally small โ except for the three systems listed above โ as derived from the efficiency being close to unity).
The orbital solutions derived in the present paper are too many to display the astrometric orbit for all cases. A few representative cases among the different subsets of Table 1 (orbital periods shorter or longer than the duration of the Hipparcos mission, small or large parallaxesโฆ) have instead been selected and are presented in Fig. 1.
### 3.1 Comparison with DMSA/O solutions
For HIP 17296 (Tc-poor S), 31205 (strong Ba) and 56731 (strong Ba), orbital solutions are provided in the DMSA/O and were derived using spectroscopic elements from the literature. HIP 50805 (dwarf Ba) is the only case in our sample where an orbit could be derived from scratch by the Hipparcos consortia. For all these systems, the astrometric orbits derived by the methods described in Sect. 2 are in excellent agreement with the DMSA/O elements, thus providing an independent check of the validity of our procedures. Further checks are presented in Sect. 3.2.
The large number (23) of systems for which orbital solutions could be extracted from the Hipparcos data (as compared to only 4 of those already present in the DMSA/O) illustrates the great potential that still resides in the Hipparcos IAD or TD.
### 3.2 Check of the astrometric orbit
Several checks are possible to evaluate the accuracy of the astrometric elements derived in the present paper.
First, it is possible to compare the astrometric and spectroscopic values of $`\omega `$, the argument of periastron. In most cases, the two determinations agree within 2$`\sigma `$ (Fig. 2). However, even when the orbital period is shorter than the Hipparcos mission, the $`\omega `$ derived from the IAD is seldom as precise as the spectroscopic one.
In a few cases, the spectroscopic orbit is assumed to be circular. In that case, the time $`T`$ of passage at periastron becomes meaningless, and is replaced by the time of the nodal passage (or, equivalently, the time of maximum radial velocity). This is equivalent to setting $`\omega `$ equal to 0. Non-zero values for $`\omega `$ would correspond to other conventions for the origin epoch. Among our systems with acceptable orbits, two have circular orbits: HIP 53763 (CH star) and HIP 52271 (strong barium star). For these systems, our fit leads to values of $`\omega `$ significantly different ($`2\sigma `$) from 0, indicating that at the reference epoch, the star is in fact far from the node where it was expected to be.
It is also possible to compare the astrometric value of $`K_1`$ (using in Eq. 12 the value of $`a`$, $`i`$ and $`\varpi `$ from the astrometry and $`e`$ and $`P`$ from the spectroscopy) with the spectroscopic value. Fig. 3 shows that, even if $`a_0/\varpi `$ is well defined, the inclinations are generally not very accurately determined, thus leading to uncertain values of $`K_1`$. This unfortunate property of $`i`$ is well illustrated on Figs. 4 and 5. One example where the accuracy of the astrometric value of $`i`$ must be questioned is the dwarf barium star HIP 105969: despite the fact that the astrometric $`a_0/\varpi `$ ratio perfectly agrees with its estimate based on the masses, the astrometric prediction of the semi-amplitude of the radial-velocity variations differs by almost of factor of 2 as compared to the actual spectroscopic value (Table 2). The only way to resolve that discrepancy is to assume that the orbital inclination is largely in error.
The semi-major axis $`R=a_0/\varpi `$ as derived from its astrophysical estimate (Eq. 13) is compared to its astrometric value in Fig. 6, and the two values are often consistent with each other. Although the value of $`a_0`$ is likely to be affected by a positive bias (i.e., a positive $`a_0`$ value is derived even when the data consist of pure noise, as clearly apparent from the astrometric $`a_0/\varpi `$ values listed in Table 1), this bias does not markedly affects the retained solutions displayed on Fig. 6, except for solutions with $`\varpi <3`$ mas, which all have $`R/\widehat{R}>1`$. Solutions for larger $`\varpi `$ values are almost equally distributed around unity.
### 3.3 Comparison TD vs IAD
Among the 81 systems studied in this paper, 22 have TD and only 5 belong to the list of accepted orbits (HIP 17296, 25092, 31205, 5080, and 105881). They hence can be used to check the consistency of the two orbit-determination methods. The astrometric orbits derived from the TD are given in Table 2 as lines labeled with โTโ in column 2. The agreement between the results obtained with the two data types is quite good (within 1$`\sigma `$ error bars). The number of TD measurements for those five systems is up to twice as large the number of IAD measurements. As a consequence, the confidence intervals of the parameters is systematically narrower. One should however keep in mind that more precise does not imply more accurate. For instance, in the case of the mild Ba star HIP 105881, the 86 TD measurements yield $`a_0/\varpi =1.59`$ (with 1$`\sigma `$ limits of 1.45 and 1.74) as compared to $`1.2\pm 0.35`$ estimated from Keplerโs third law. The 54 IAD measurements from FAST and NDAC combined yield a more accurate result of 0.99 (between 0.44 and 1.69).
For the sake of completeness, we checked the astrophysical consistency of the orbit derived for the 22 systems for which TD are available. All systems but the five already accepted were rejected. In essence, this confirms that the astrometric content of the TD is basically equivalent to the IAD and that nothing new can come out from TD if IAD do not yield a reliable solution.
## 4 Comparison of parallaxes and proper motions derived from single- or double-star solutions
It has been stressed (Wielen (1997); Wielen et al. (1999)) that the proper motions provided by the Hipparcos catalogue may be systematically in error for long-period binaries (i.e. with $`P>3`$ yr), if those were not recognized as such by the reduction consortia. The orbital motion may then add to the actual proper motion, changing both its direction and modulus. The present sample offers a good opportunity to evaluate the impact of this effect.
Figs. 7 and 8 compare the position angle and modulus, respectively, of the Hipparcos proper motion with those derived when account is made of the orbital motion as in the present work. It is clearly apparent that the direction of the proper motion listed in the Hipparcos catalogue is correct, as expected, for orbits with periods less than 3 yr, corresponding to the duration of the Hipparcos mission. The Hipparcos values are the less accurate in the orbital-period range 3 to about 5 yr. The position angle quoted by the Hipparcos catalogue may then be off by several dozens degrees (a good illustration of this situation is offered by the the strong Ba star HIP 110108 in Table 2), though it differs generally by less than $`3\sigma `$ from the correct value. For longer orbital periods, the orbital motion becomes negligible over the duration of the Hipparcos mission, and the proper motions are again rather well determined in the Hipparcos catalogue. The situation is almost identical for the the proper-motion modulus, except that the error bars now become very large for orbital periods longer than 3 yr. This situation translates the fact that the proper motion and the semi-major axis are strongly correlated (resulting in a large formal uncertainty on the proper motion), because the two motions are difficult to disentangle when the Hipparcos data sampled only a fraction of the orbit. This situation is encountered e.g. for HIP 104785 and 104732 (whose astrometric orbits have not been retained precisely because of the large uncertainty on $`a_0`$ introduced by the strong correlation with $`\mu `$), and results in sometime large differences between the $`a_0`$ and $`\mu `$ values derived from the NDAC, FAST and NDAC+FAST data sets, since they are now very sensitive to the measurement errors.
Fig. 9 compares the parallax listed in the Hipparcos catalogue with that derived in the present work, as a function of the orbital period. It turns out that the agreement is good, except for systems with orbital periods close to 1 yr. For those systems, the parallax cannot be accurately derived since the orbital motion and the parallactic motion are strongly entangled. This situation is encountered for HIP 17402, 53763 and 62827. Large error bars at periods different from 1 yr correspond to systems with parallaxes smaller than 3 mas.
A few stars (HIP 8876, 29740, 29099, 32831, 36613) in our sample had negative parallaxes in the Hipparcos catalogue. The parallaxes we derive for these systems are slightly positive ($`\varpi ^{}`$ was adopted to guarantee that property), but the associated error bar encompasses zero, so that these new parallaxes are just useless.
## 5 Masses
The test presented in Sect. 2.1.5 (Eq. 13) to evaluate the astrophysical relevance of the orbital solution on the basis of $`a_0/\varpi `$ is in essence based on an a priori knowledge of the masses, since Eq. 13 may in fact be rewritten as
$$Q=\frac{M_2^3}{(M_1+M_2)^2}=\frac{(a_0/\varpi )^3}{P^2}.$$
The orbital solutions that appear to be statistically significant on the basis of the F-test in almost all cases turned out to have a semi-major axis $`a_0/\varpi `$ consistent with its expected value based on the estimated masses, the dwarf barium star HIP 60299 being however a notable exception. This agreement is well illustrated on Fig. 6.
Therefore, the present analysis does not add much to our previous knowledge of the masses, especially since the astrometric orbit only allows to eliminate $`i`$ from the mass function but does not give access to the individual masses.
It was originally hoped that the present astrometric results might be used to test the hypothesis which played a central role in the statistical analysis of the mass functions of CPRS (Jorissen et al. (1998)), namely that their $`Q`$ distributions are quite peaked, because they host a white dwarf companion. However, the error bar on $`a_0/\varpi `$ (and hence on $`Q`$) is too large, even for the best determined barium-dwarf orbits (Table 2), to be able to draw a meaningful $`Q`$-distribution from the astrometric orbits.
## 6 Conclusions
The availability of the Hipparcos Intermediate Astrometric Data and, when they exist, the Transit Data, allow to improve upon the existing Hipparcos catalogue. Once a new or revised spectroscopic binary orbit becomes available, it can be used to update the Hipparcos astrometric parameters and, sometime, to derive the corresponding astrometric orbit. In the latter case, the orbital inclination is obtained, thus allowing to improve our knowledge of the binary system.
The optimization method used in the present paper yields a solution in all cases, and it is therefore important to perform consistency checks. One test which may be applied in all cases involves the comparison of the arguments of the periastron and of the semi-amplitude of the radial-velocity curve, as derived from either the spectroscopic orbit or the astrometric one. The comparison of the semi-amplitude of the radial-velocity curve involves however the inclination of the orbit which is not always very well determined.
In this context, chemically-peculiar red stars (CPRS) like barium, Tc-poor S and CH stars are interesting because they provide their own consistency checks. Indeed, beside the orbit which can be determined as for any other spectroscopic binary, the mass of both components may be guessed with some confidence. These mass estimates based on astrophysical considerations may then be compared to the mass of the system derived from the astrometric orbit using Keplerโs third law.
From a sample of 81 CPRS spectroscopic systems whose orbits became available after completion of the Hipparcos catalogue, we have derived 23 reliable astrometric orbits. This shows that the number of Hipparcos entries for which an orbital solution may be obtained is much larger than suggested by the number of existing entries in the DMSA/O.
Updated astrometric parameters from this particular sample have shown that the Hipparcos-catalogue parallaxes are not reliable for systems with orbital periods close to 1 yr. Similarly, the Hipparcos proper motions are not reliable for unrecognized binaries with periods in the range 3 to about 8 yr, as already suspected by Wielen et al. (1999).
The orbit-determination methods based on the IAD and TD being completely different, systems for which the two sets of data are available can be used to assess the consistency of the solutions derived from both sets with different methods. It turns out that the two sets of results are generally in good agreement.
###### Acknowledgements.
We thank P. North for communicating us the orbital elements of dwarf Ba stars in advance of publication. We also thank two anonymous referees for their very constructive comments.
|
warning/0006/astro-ph0006435.html
|
ar5iv
|
text
|
# DISTRIBUTION OF DUST FROM KUIPER BELT OBJECTS
## 1. Introduction
It is well known that the Edgeworth-Kuiper belt objects (called โkuiperoidsโ hereinafter) can replenish the cometary populations throughout the Solar system (e.g., Levison & Duncan 1997; Ozernoy, Gorkavyi, & Taidakova 2000a,b $``$ OGT 2000a,b; and refs. therein). Recently, it has been recognized that kuiperoids might be also one of the major sources of dust in the Solar system (e.g., Backman et al. 1995, Liou et al. 1996). This dust can be produced due to evaporation of the volatile material from the surface of kuiperoids by processes, such as the solar radiation and wind, mutual collisions of kuiperoids, micrometeor bombardment, etc. In this paper, we perform extensive numerical simulations to examine the distributions (both in orbital parameters and in space) of kuiperoidal dust particles and thereby to analyse the structure of the kuiperoidal dust cloud.
The dynamics of this dust is determined by three major effects: (i) the Poynting-Robertson (P-R) drag (including radiation pressure and solar wind drag), (ii) gravitational scattering on the planets, and (iii) resonances with the planets. Extensive work on dust particle evolution governed by these effects has been done by a number of investigators (Weidenschilling & Jackson 1993; Hamilton 1994; Roques et al. 1994; Liou & Zook 1997; Gorkavyi, Ozernoy, Mather, & Taidakova ($``$ GOMT) 1997a,b and 1998a,b; Kortenkamp & Dermott 1998). The evolution of kuiperoidal dust was analyzed by Liou, Zook, & Dermott (1997), Gorkavyi, Ozernoy, & Taidakova (1998), and Liou & Zook (1999). The present paper makes a next step by employing a fast and efficient method to compute a stationary distribution function of dust particles in the phase space, which provides much better statistics to derive a 3-D model of the interplanetary dust cloud and to reveal its rich resonant structure.
In Sect. 2, we discuss the sources of dust in the outer Solar system. The dynamical evolution of dust particles is reviewed in Sect. 3. Sect. 4 describes our numerical method that enables us to compute the 3-D distribution of dust in the Solar system. We employ an implicit second-order integrator for dissipative systems (Taidakova & Gorkavyi 1999) outlined in Appendix. Sect. 5 contains the results of these computations, which reveal the global dust distribution as well as interesting details of its resonant structure. Our conclusions are presented in Sect. 6.
## 2. The Sources of Dust Particles
There is mounting evidence that the sources of the interplanetary dust particles (IDPs) cannot be entirely reduced simply to those comets which produce the observed dust tails and/or to asteroids which are thought to be responsible for the observed โdust bandsโ in the IDP emission. Many facts forces us to suspect that additional sources of the interplanetary dust must exist. Among others, two groups of facts are worth mentioning: (i) According to Pioneers 10 & 11 and Voyagers 1 & 2 data, the dust number density is approximately distance-independent in the outer Solar system between 10 and 40-50 AU (Humes 1980, Divine 1993, Gurnett et al. 1997), while both the asteroidal and cometary dust number densities are known to sharply decrease with heliocentric distance (GOMT 1997b); (ii) Chemical analyses and other space-based data indicate that some IDPs spent a much larger time in space than the typical asteroidal and cometary particles (Flynn 1996). Thus, there is strong evidence in favor of other sources of dust in the Solar system, along with the known comets and asteroids.
This third component of the IDP cloud might be the โkuiperoidalโ dust (Backman et al. 1995). The Kuiper belt can influence the formation of the IDP cloud in two ways: 1) as a source of small-size particles slowly drifting toward the Sun under the combined action of the PR-drag, gravitational scattering, and resonances; and 2) as a source of trans-Jovian comets. It is commonly agreed that the Jupiter-family comets are produced by transporting the comets from the Kuiper belt via gravitational scattering on the four giant planets so that each planet scatters the comets both toward and away from the Sun (Levison & Duncan 1997). Our simulations (OGT 2000a,b) indicate that, between Jupiter and Neptune, there is a large population of minor bodies forming four cometary-asteroidal belts near the orbits of the giant planets.
According to our simulations, the minor body families of Saturn, Uranus, and Neptune should contain progressively larger numbers of comets than one sees near Jupiter. Even despite a many-fold decrease of the solar heat intensity at such large distances, those numerous comets may produce dust in amounts comparable to that from a few active J-comets. Complementary mechanisms of dust release from kuiperoids and Centaurs between Jupiter and Neptune can include impacts of large grains and the solar wind. Without discussing here the dust production by the above mechanisms, which is out of our paperโs scope, we simply refer to observational data that indicate, for a number of kuiperoids and Centaurs, a steady cometary activity lasting for years (e.g. Brown & Luu 1998 and refs. therein).
## 3. Dynamical evolution of dust particles
The kuiperoidal dust experiences the same dynamical effects as the cometary and asteroidal dust, with the only difference that, due to a slower PR-drift and a stronger influence of the giant planets, the role of gravitational scattering and resonance captures must be more important for it.
Just after the birth of a dust particle, the solar pressure forces it to change its orbit to a more distant and eccentric one. This change of the orbital parameters (semimajor axis $`a_d`$ and eccentricity $`e_d`$) of dust particles, which start their journey from the apocenter or pericenter of kuiperoidal orbits, is described by
$$a_d=a_K\frac{1\beta }{12\beta \left(1\pm e_K\right)^1},$$
$`(1)`$
$$e_d=\left(1\frac{(1e_K^2)\left[12\beta \left(1\pm e_K\right)^1\right]}{(1\beta )^2}\right)^{1/2}.$$
$`(2)`$
Here $`a_K`$ and $`e_K`$ are the orbital parameters of kuiperoids, signs โ$`+`$โ and โ$``$โ correspond to the start from apocenter and pericenter, accordingly, and $`\beta L_{}/(M_{}r)`$ is the ratio of the solar light pressure and the gravitational force applied to the dust grain of radius $`r`$ (in $`\mu `$m). In what follows, we consider two $`\beta `$-values: $`\beta =0.285`$ and $`\beta =0.057`$, which correspond, for grains of density $`\rho =(12)`$ g/cm<sup>3</sup>, to $`r=(12)\mu `$m and $`r=(510)\mu `$m, respectively.
The basic mechanisms of the dynamical evolution of dust are as follows:
1. Drift of particles toward the Sun under the Poynting-Robertson drag. Such drift occurs more or less freely except for short periods when particles experience a strong gravitational scattering on the planets or when they pass through the outer or inner resonances without being captured into them. A particle governed by the P-R drag drifts in the $`(a,e,i)`$-space, with a decreasing $`a`$ and $`e`$, along the line
$$a(1e^2)e^{4/5}=\mathrm{const},i=\mathrm{const}.$$
$`(3)`$
2. Gravitational scattering by the planets, which builds up a population of particles with large eccentricities, including grains that are ejected from the Solar system. A gravitationally scattered particle experiences a rather chaotic โjumpโ in the $`(a,e,i)`$-space, but conserves its Tisserand parameter $`๐ฏ`$ during the jump:
$$๐ฏ=\frac{1}{2}\left(\frac{a_{pl}}{a}\right)+\left(\frac{a}{a_{pl}}\right)^{1/2}\left(1e^2\right)^{1/2}\mathrm{cos}i=\mathrm{const}.$$
$`(4)`$
Although the P-R drag changes that parameter, it happens on a much longer timescale. As shown in OGT (2000b), gravitational scattering results in the motion of most of the particles within the so called โcrossing zoneโ (i.e. the zone of strong gravitational scattering) defined, in the $`(a,e)`$-plane of orbital coordinates, by:
$$a(1e)a_{pl}\mathrm{if}a>a_{pl},$$
$$a(1+e)a_{pl}\mathrm{if}a<a_{pl}.$$
$`(5)`$
Here $`a`$ and $`e`$ are the semimajor axis and eccentricity of the test particle, respectively; and $`a_{pl}`$ is the semi-major axis of the planet. As a result of gravitational scatterings, the particle pericenters are close enough to the planetโs orbit $`a_{pl}`$ (OGT 2000b):
$$a(1e)a_{pl}\mathrm{if}a>a_{pl}.$$
$`(6)`$
3. Resonant capture of dust into outer resonances with the planets. A particle captured into a resonance is positioned on the line
$$a=(1\beta )^{1/3}\left(\frac{n}{m}\right)^{2/3}a_{pl},$$
$`(7)`$
where usually $`n>m`$ (i.e. this is an outer resonance). The eccentricity of the resonant particle increases with time (Weidenschilling and Jackson 1993), while its $`i`$ oscillates and gradually decreases (Liou & Zook 1997).
Athough the above dynamical trajectories governed by just one dominating dynamical factor turn out to be only approximate as soon as the two other factors are taken into account, they are very helpful as analytical approximations in the appropriate limiting cases. Moreover, in accordance with the dominating dynamical factor, we expect to get three major components of dust populations: i) โfreelyโ drifting particles, (ii) gravitationally scattered particles, and (iii) particles captured into resonances. This classification is helpful while interpreting the results of the present numerical simulations. In Sect. 5, we consider the steady-state distribution of these three dust components, but before that we describe our computational approach.
## 4. Computational Method: Simulation of a Quasi-stationary Distribution of Dust Particles in the Solar System
We calculate the orbital elements $`(a,e,i)`$ and then the spatial positions of massless particles starting from a particular kuiperoid as the source of dust and drifting toward the Sun under the P-R drag. On its way to the Sun, each particle undergoes the gravitational influence of the planets. To save computational effort, we assume that the Sun is fixed at the origin and the 8 planets (excluding Pluto) are on circular orbits with zero inclinations (this approximation will be abandoned in our further work).
The distribution of dust in the Solar system averaged over a time scale of planetary orbital motions is described as stationary. To simulate this stationary distribution of dust particles, we applied the following procedure: We computed the dynamical trajectory of each particle and kept a record of the particleโs orbital elements and positions with a certain time interval. These data were then used to characterize the orbital elements and positions of many particles over the entire time span, from an initial instant intil the particleโs death (impact on planet, the Sun, or particleโs ejection from the Solar system). For a stationary system, such as the Sun, the constant number of dust grains (being in a balance between their production by minor bodies and eventual disappearance in due course of a drift toward the Sun under the P-R drag), and no planets included, this approach would not require a detailed proof. For a system that incorporates several planets, any particular grainโs trajectory cannot be considered as a stationary one, the cloning of one dust particle into many others as outlined above is justified by the following arguments.
After a dust particle starts its journey, gravitational scattering on the planets causes chaotic change of the particleโs orbital elements. For example, we found that the orbital parameters of a particle of $`r=5`$-$`10\mu `$m changed by $`\genfrac{}{}{0pt}{}{_>}{^{}}1`$% in $`10^3`$ yrs and by $`\genfrac{}{}{0pt}{}{_>}{^{}}10`$% in $`10^5`$ yrs. In other words, during the particleโs lifetime of $`210^7`$ yrs, there are hundreds of strong changes in its trajectory. Due just to gravitational scatterings on the giant planets, the scattered particles rapidly forget their initial conditions (Levison & Duncan 1997). Unless the number of particle trajectories is very small, the computed distribution function of dust particles depends only weakly on initial $`(a,e,i)`$-orbital elements for the particular trajectory, and practically does not depend on the time the trajectory starts. Therefore, in deriving the particle distribution function for such a highly chaotic system as the dust population in the outer Solar system, the initial positions of kuiperoids as the dust sources are much more important than the start time for particles.
For the present paper, we simulated 280 trajectories of dust particles starting from the apocenter and pericenter of 100 KBOs, which produced $`1.2\times 10^{11}`$ particle positions. In the course of our computations of each particleโs trajectory, the following procedure was applied:
1. The computed particle positions were sorted on a 3D-grid containing $`45\times 180\times 244210^6`$ cells with steps in (heliocentric latitude $`\phi `$, longitude $`\lambda `$, and radius $`R`$) of ($`2^{},2^{},0.025R`$ \[AU\]).
2. The computer recorded the particleโs $`(a,e,i)`$ orbital elements and $`(x,y,z)`$-coordinates once per revolution of Neptune around the Sun (i.e. every 164.8 years). This enabled us to create an auxillary file containing $`1.210^{11}/610^3=210^7`$ particle positions. These coordinates were sorted into two 2D data files: a $`100\times 1000`$ array in the $`(a,e)`$-plane ($`a<150`$ AU) and a $`180\times 1000`$ array in the $`(a,i)`$-space. The following bins were used: $`\mathrm{\Delta }a=0.15`$ AU, $`\mathrm{\Delta }e=0.01,\mathrm{\Delta }i=0.5^{}`$. A few auxiliary 1D files dealing with particle distributions in semimajor axis, $`n(a)`$, in pericentric distance, $`n(q)`$, and in radius, $`n(r)`$, all used $`\mathrm{\Delta }=0.3`$ AU.
Let us estimate the number of trajectories which would be sufficient to derive the stationary distribution function. Suppose that every cell of the $`(a,e,i)`$-space or $`(x,y,z)`$-space is permeated by $`N`$ trajectories. The necessary condition to get a robust distribution of particles in any cell is that a very large number of trajectories, $`N>>1`$, visited that cell. For practical reasons, we adopt $`N10`$. Since the trajectories are highly chaotic, it does not matter whether a particular cell is visited, say, 10 times by the same particle or by 10 different particles just once. We found that a particle of $`r=5`$-$`10\mu `$m requires about $`210^3`$ yrs (on average) to change one $`(a,e)`$-space cell for another, so that it visits $`10^4`$ cells during its lifetime. Therefore, on a 2D $`(a,e)`$-grid containing $`100\times 1000=10^5`$ cells, one needs to simulate $`100`$ different trajectories to achieve, on average, $`N10`$ particle visits per cell (actually, this number may be as large as a few hundred in the densest regions and $`10`$ in the rarefied ones). For the present paper, we similated 80 trajectories of $`r=5`$-$`10\mu `$m particles and 200 trajectories of $`r=5`$-$`10\mu `$m particles, which provides a resonable statictics with a rather low level of noise.
We employ a second-order implicit numerical integrator described in Taidakova (1997) and Taidakova & Gorkavyi (1999). Its basic features are outlined in the Appendix. As shown there, for dust particles produced by typical kuiperoids the dissipative integral of motion is conserved with an accuracy of $`10^410^3`$, which is appropriate to explore such a highly chaotic system as dust population in the Solar system.
## 5. The Results: Dust Belts and Their Resonant Structure
Using the above computational approach, we have simulated the spatial structure of the IPD cloud between $`0.5`$ and $`100`$ AU, and determined distributions in the phase space of orbital elements. Here we present the results for dust particles of radius 1-2 $`\mu `$m produced by 100 Kuiper belt objects, and particles of radius 5-10 $`\mu `$m produced by 40 KBOs. The initial conditions for orbit integration were taken in accordance with equations (1)-(2).
The distributions have been computed with two different values of the P-R parameter (the radiation pressure to gravitational force ratio) $`\beta =0.057`$ and 0.285, and with the solar wind drag to the PR-drag ratio $`=`$0.35 (Gustafson 1994). Details of computational runs are given in Table 1. The total CPU time of those computations with a 450 MHz PC was about 2 months.
### 5.1. Distribution of Dust in Phase Space
Distribution of Dust in $`(a,e)`$\- and $`(a,i)`$-space. A representative dynamical trajectory of a dust particle on the phase plane is shown in Fig. 1, where the changing role of different forces at different parts of the trajectory is clearly seen. The trajectory is ended with the ejection of the particle from the Solar system by Jupiter, which is typical for most particles. The rest of particles eventually find their death on the the Sun or the planets. Our computations indicate that about 11.5% of $`1`$-2$`\mu `$m particles penetrate the innermost zone of the Solar system, whereas 88.5% of particles are ejected. Similarly, 13.8% of $`5`$-10$`\mu `$m particles penetrate the Earth zone, whereas 85% are ejected from the Solar system (and one of the 80 particles, i.e. 1.2%, fell onto Neptune). This $`12`$% fraction of kuiperoidal particles penetrating the inner Solar system is consistent with what was found by Liou, Zook, & Dermott (1997). This fraction is not very sensitive to the particle size, which can be explained as followss: although smaller particles drift more quickly toward the Sun under the P-R drag, they are more easily ejected from the Solar system due to the combined action of the solar radiative pressure and gravitational scattering by the planets.
The overall picture of dust distribution in the Solar system obtained by summation of the computed trajectories is shown in Figs. 2a,b using 80 trajectories of large ($`r=5`$-10$`\mu `$m) particles and in Figs. 2c,d using 200 trajectories of small ($`r=1`$-2$`\mu `$m) particles. We find the simulated dust distribution be highly non-uniform. All three dynamical classes of dust particles listed at the end of Sect. 3 are clearly seen on the $`(a,e`$)-plane of Fig. 2c:
โ drift particles with a maximum density at $`a>45`$ AU for $`1`$-2$`\mu `$m particles and at $`a>30`$ AU for $`5`$-10$`\mu `$m particles;
โ resonant particles producing numerous dense populations in the Neptunian and other resonances; the resonant population of large particles has a bigger contrast with the background than the population of small particles;
\- scattered particles stretched along the right boundary of the planet crossing zone (i.e. the particle pericenters are located near the planetโs orbit), especially for Neptune, Saturn, and Jupiter. Scattered particles tend to avoid resonant orbits thereby forming resonant gaps. The Neptunian region possesses the most dense population of scattered particles.
The dust distribution in the $`(a,e)`$-space shown in Figs. 2a,c indicates that each planet governs a dust belt that consists of both resonant and scattered particles. If the resonant particles dominate in the belt, then the latter is associated with a spatial excess of dust particles (like the Neptunian belt described in Sect. 5.2). A different situation arises if the belt consists mostly of scattered particles: in this case, the belt is characterized by a density minimum (like the Jovian belt). In both cases, the belt is revealed and can be seen as a maximum in the particle distribution in distance of pericenter, as will be shown at the end of this Section. A criterion for a dust particle to be included into the belt is a substantial dynamical interaction of the particle with the host planet, either by resonance or by gravitational scattering.
As can be seen from the dust particle distribution on the $`(a,i)`$-plane, the inclinations of kuiperoidal particles, on average, substantially increase due to gravitational scattering as the particles move from the Kuiper belt toward Jupiter.
Distribution of Dust Particles in Semimajor Axis and Their Resonant Structure. Fig. 3 shows the distribution of kuiperoidal dust in semimajor axis. The vertical coordinate is a measure of the number of particles within each 0.3 AU bin per trajectory. Fig. 3 reveals a rich resonant structure of each dust belt. Arrows show positions of particular resonances. One can see that large-size particles are more easily captured in (or spend more time within) the resonances than small-size particles. Furthermore, the smaller the particle size, the smaller is the contrast of the resonant structure to the background, which confirms a similar conclusion reached by Liou & Zook (1999). Large particles with $`a>50`$ AU are mostly scattered particles, therefore one can see resonant gaps in their distribution rather than resonant peaks, whereas peaks are seen at smaller $`a=3550`$ AU. Indeed, as Fig. 3 demonstrates, the fraction of resonant particles in the Neptunian dust belt is very large.
The following resonant features seen in Fig. 3 are worth mentioning:
in the scattered dust component, the gaps at the resonances 5:2, 7:2, 4:1, 5:1, 6:1 etc., are pronounced;
for captured particles, peaks at the resonances 6:5, 4:3, 3:2, 5:3, 7:4, 9:5, 2:1, 4:1, 5:1, 6:1, etc., are pronounced;
Like the cometary belts simulated in OGT (2000a,b), the simulated dust belts indicate a complex structure containing many families of captured resonant particles and gaps. While resonances in the cometary population can form in a non-dissipative way, the resonant capture of dust particles occurs dissipatively, and this process takes place both inside and outside the crossing zone. For $`\beta 0`$, resonances are shifted by a factor $`(1\beta )^{1/3}`$. The larger the value of $`\beta `$, the larger is the drift velocity and the smaller is the probability of a resonant capture.
In a resonance $`(j+1)/j`$, while the eccentricity is close to the maximum, $`e_{\mathrm{max}}=\sqrt{0.4/(j+1)}`$ (Weidenschilling & Jackson 1993), the particleโs resonant lifetime is expected to be long. The resonances seen in Figs. 2a and 2c demonstrate this kind of behaviour: the larger is eccentricity of particles in a given resonance, the more abundant is their population. This results in a two-hump structure seen, for large-size particle distribution, between $`q25`$ and 40 AU in Fig. 4 and between $`R25`$ and 45 AU in Fig. 5. The characteristic shape of such dust structures follows from the Kepler laws (Kessler 1981, GOMT 1997b): The inner and outer edges of each structure are given by $`a_{res}(1e_{\mathrm{max}})`$ and $`a_{res}(1+e_{\mathrm{max}})`$, respectively. For all resonances, in a good approximation, the position of the inner edge is $`0.85a_{\mathrm{planet}}`$. At this position, we expect to find rather sharp inner edges and steps in the dust density distribution (somewhat depending on the particle size) for each giant planetโs dust belt.
Direct measurements of the expected dust density maximum at $`R2729`$ AU by dust detectors on spacecraft would be a strong confirmation of the predicted Neptunian dust belt, as well as its resonant nature simulated in the present work.
Distribution of Dust Particles in Pericenter Distance. Fig. 4 demonstrates that the pericentric distances of many kuiperoidal particles are located close to the orbit of each giant planet (see also Figs. 2a,c). One important difference between the resonant and the scattered components of a belt is in the value of distance of pericenter, which is $`q_r(0.851.0)a_{pl}`$ for the resonant component (see the location of the Neptunian belt relative to the Neptuneโs orbit shown as N in Fig. 4), and $`q_s(1.01.1)a_{pl}`$ for the scattered component (see the locations of the Jovian and Saturnian belts relative to the respective planets shown as J and S). Thus the distribution in pericentric distance provides an additional strong argument in favor of the four dust belts. Earlier (OGT 2000a,b), we have found that cometary bodies concentrate into belts near each giant planetโs orbit. As the parameter $`\beta `$ decreases with increase of particle size, the efficiency of gravitational scattering increases, which makes the pericentric distribution of large dust grains like that of comets.
### 5.2. Spatial Distribution of Kuiperoidal Dust Particles
Fig. 5 presents the column density of dust particles \[using all recorded particle positions $`(x,y,z)`$\] as a function of heliocentric distance. Going inward, the following features of dust density distribution are worth emphasizing:
an approximately constant column density of dust (as well as volume density) between 50 and 10 AU;
a sharp decrease of both column and volume density of dust between 10 and 4 AU, which is due to the ejection of particles by Jupiter and Saturn;
an increase of column density of dust (accompanying by an even steeper increase of volume density near the ecliptic plane) at $`R<4`$ AU.
The column density of kuiperoidal dust forms a plateau between 50 and 10 AU, and this seems to be the most remarkable result. For 5-10 $`\mu `$m dust, the value of column density is several times higher than that for 1-2 $`\mu `$m grains, which is explained by a slower rate of evolution of larger size particles, mostly due to the P-R drag. The influence of gravitational scattering and resonances is also different for particles of different sizes, so that the shape of the plateau, as can be seen from Fig. 5, depends upon particle size.
The distribution of the number density of dust in the ecliptic plane, which is of obvious practical interest to interpret the data of space missions, is qualitatively similar to column density graph. This is because the particle inclination, on average, decreases outward (see Figs. 2b,d). A region of an elevated density of kuiperoidal dust, which is associated with the Neptuneโs orbits, can be seen in Fig. 5. In the ecliptic plane, the Neptunian dust belt is expected to have the largest number density of particles. The major part of the simulated Neptuneโs dust belt consisting mostly of resonant particles is located between 25 and 45 AU and forms a flat dense disk. The simulated Uranian, Saturnian, and Jovian dust belts basically overlap and form complex structures which, in their central parts, are less dense than the Neptunian dust belt.
There is a remarkable density minimum between Mars and Jupiter. This minimum is due to the fact that Jupiter either ejects from the Solar system or transfers to more inclined and eccentric orbits an appreciable part of the dust drifting toward the Sun. An increase of dust number density going inward from Mars to Earth can be explained by the dominant role of the P-R drag here, which would result in the number density distribution $`n(R)R^1`$ for circular orbits or $`n(R)R^{(23)}`$ for eccentric orbits (see GOMT 1997b and refs. therein).
Fig. 6 shows a 2D section of the spatial structure of kuiperoidal dust number density in the region up to 10 AU perpendicular to the ecliptic plane (edge-on view). An important feature found in our simulations is the existence of a new quasi-stationary, highly inclined dust population with pericenters near Jupiter and Saturn. This population is seen in Fig. 6 as a sharp increase in dust density beyond the Jupiter orbit, which looks like a โChinese wallโ. This structure is found to be steeper and denser for larger particles. The quasi-stationarity of the structure results from a balance between the tendencies for particleโs semimajor axis $`a`$ and eccentricity $`e`$ to increase due to gravitational scattering on the planet and to decrease due to the P-R drag. The particle inclinations increase substantially due to gravitational scattering and resonances.
## 6. Discussion and Conclusions
We have explored whether the 280 particle trajectories used in the present study are sufficient to provide reliable results. As can be seen from Fig. 5, the use of just 20 particle trajectories already reveals the major features of large-size particle distribution, but for increased accuracy we used 80 trajectories. To explore the detailed distribution of kuiperoidal dust in phase space, we need as many as 80 trajectories to reveal the basic resonant features (see Figs. 2a,b). The comparison of Figs. 2a,b with Figs. 2c,d (where 200 trajectories are used) indicates that a larger number of computed trajectories leads to a decrease in random fluctuations; gives a clearer picture of strong resonances free of discontinuities; reveals a larger number of weak resonances; and results in a more reliable picture of resonant gaps and near-resonant accumulations in the scattered component of dust. Although even larger statistics would certainly provide further improvements, we do not think that it would qualitatively change the results.
This study employs a number of approximations and simplifications (circular planetary orbits with zero inclinations, neglecting other components of dust, including the interstellar dust), which will be treated more accurately in further work. Flynn (1996) and Liou et al. (1997) claim that large ($`\genfrac{}{}{0pt}{}{_>}{^{}}9\mu `$m) kuiperoidal dust is destroyed by collisions with interstellar dust. Nevertheless, it is instructive to compare our results with available observational data.
Pioneers 10 and 11 as well as Voyagers 1 and 2 detected a large number of dust particles in the region between Jupiter and Neptune (Humes 1980; Gurnett et al. 1997). These data show that the number density of dust particles in the outer Solar system is approximately distance-independent. As can be seen in Fig. 5, the simulated distribution of kuiperoidal dust can explain this constancy. This qualitative result will be quantified in our further work, after improving our modelling and taking into account relative velocities of dust particles and the spacecraft.
The approximate constancy of the dust density between 10 and 50 AU found in the our simulations differentiates the kuiperoidal dust from both the asteroidal and cometary dust, whose number densities are known to appreciably decrease with distance from the Sun (for interpretation, see GOMT 1997b and refs. therein).
Our simulations offer a high-resolution, 3-D model of the kuiperoidal cloud on a grid of 2 million cells containing 115 billion computed positions of dust particles. The major conclusion reached in the present simulations are as follows:
1. The simulated dust distribution is highly non-uniform. Moving inward, the column and volume density of kuiperoidal dust is approximately constant at heliocentric distances from 50 to 10 AU, sharply decreases between 10 and 4 AU giving a deep minimum, after which the column density of dust increases (accompanied by a constant or decreasing volume density near the ecliptic plane) at $`R<4`$ AU.
2. We find a new quasi-stationary, highly inclined dust population with pericenters near all giant planets. This quasi-stationarity results from a balance between the tendencies for the particle semimajor axis $`a`$ and eccentricity $`e`$ to increase due to gravitational scattering on the planet and to decrease due to the P-R drag. The particle inclinations $`i`$ increase substantially due to gravitational perturbations from Jupiter and Saturn.
3. Most of the dust is concentrated into four belts consisting of resonant and scattered particles associated with the orbits of the four giant planets. These belts are chiefly of dynamical nature. Some of them, such as the Neptunian belt, disclose first of all as dust structures. The others have very distinct distributions in the phase space: for instance, the Jovian (Saturnian) belt is characterized by a substantial excess of highly eccentric particles with the maximum of dust distribution in pericentric distance at $`q5(10)`$ AU. The Saturnian belt could be discriminated by an excess of dust particles in the resonances 4:3, 3:2, and 2:1.
4. The simulated dust belts reveal a complex resonant structure containing many families of gaps and resonant maxima. The particles are either dissipatively captured into exteriour resonances (usually outside the crossing zone) or form gaps (usually inside the crossing zone).
5. A rather long life time in each resonance, while the eccentricity is close to the maximal one, results in a steep rise of dust density at the innermost edge of the resonant component in all dust belts, especially the Neptunian one. Rather sharp inner edges and โstepsโ in the dust density distribution are expected to characterize each giant planetโs dust belt at 0.85$`a_{\mathrm{pl}}`$, where $`a_{\mathrm{pl}}`$ is the planetโs semimajor axis. Neptuneโs dust belt is expected to have both the largest โstepโ and number density of particles in the ecliptic plane (Fig. 5). Direct detection of a dust density maximum at $`R2729`$ AU in Neptuneโs zone would test the simulated dust distribution. The Neptunian and the other belts would be a challenging target to discover and measure by space missions.
The resonant features of dust distributions near giant planets can serve as signatures of exo-planets in the circumstellar disks (Ozernoy et al. 2000c, Gorkavyi et al. 2000a). The kuiperoidal dust is likely to be a contributor of the zodiacal light emission in the Solar system, which is analyzed in Gorkavyi et al. (2000b).
Acknowledgements. This work has been supported by NASA Grant NAG5-7065 to George Mason University. N.G. acknowledges the NRC/NAS associateship.
## Appendix
To simulate the dissipationless dynamics of comets as well as dissipative dynamics of dust particles orbiting around a star, we use an implicit second-order integrator (Potter, 1973; Taidakova, 1997; Taidakova & Gorkavyi, 1999; Fridman & Gorkavyi, 1999) in a rotating (comoving with Neptune) coordinate system. The latter is convenient to show the planetary resonant structure as a stationary pattern.
The equations of motion of a dust particle in the gravitational field of the Sun and the planets in this coordinate system take the form (Taidakova 1990, 1997):
$`\ddot{x}`$ $`=`$ $`2\dot{y}+x+F_x`$
$`\ddot{y}`$ $`=`$ $`2\dot{x}+y+F_y`$
$`\ddot{z}`$ $`=`$ $`F_z,`$
where $`F_x,F_y,F_z`$ are the components of the sum of the gravitational forces and the PR-drag. Integration employs the following equations (Taidakova 1990, 1997):
$`v_x^{[n+1]}`$ $`=`$ $`{\displaystyle \frac{v_x^{[n]}(1\mathrm{\Delta }^2t)+(2v_y^{[n]}+x^{[n+\frac{1}{2}]}+F_x^{[n+\frac{1}{2}]})\mathrm{\Delta }t+(y^{[n+\frac{1}{2}]}+F_y^{[n+\frac{1}{2}]})\mathrm{\Delta }^2t}{1+\mathrm{\Delta }^2t}}`$
$`v_y^{[n+1]}`$ $`=`$ $`{\displaystyle \frac{v_y^{[n]}(1\mathrm{\Delta }^2t)(2v_x^{[n]}y^{[n+\frac{1}{2}]}F_y^{[n+\frac{1}{2}]})\mathrm{\Delta }t+(x^{[n+\frac{1}{2}]}+F_x^{[n+\frac{1}{2}]})\mathrm{\Delta }^2t}{1+\mathrm{\Delta }^2t}}`$
$`v_z^{[n+1]}`$ $`=`$ $`v_z^{[n]}+F_z^{[n+\frac{1}{2}]}\mathrm{\Delta }t`$
$`x^{[n+1]}`$ $`=`$ $`x^{[n]}+(v_x^{[n+1]}+v_x^{[n]})\mathrm{\Delta }t/2`$
$`y^{[n+1]}`$ $`=`$ $`y^{[n]}+(v_y^{[n+1]}+v_y^{[n]})\mathrm{\Delta }t/2`$
$`z^{[n+1]}`$ $`=`$ $`z^{[n]}+(v_z^{[n+1]}+v_z^{[n]})\mathrm{\Delta }t/2,`$
where $`x^{[n+\frac{1}{2}]}=x^{[n]}+v_x^{[n]}\mathrm{\Delta }t/2;y^{[n+\frac{1}{2}]}=y^{[n]}+v_y^{[n]}\mathrm{\Delta }t/2`$ ; $`z^{[n+\frac{1}{2}]}=z^{[n]}+v_z^{[n]}\mathrm{\Delta }t/2;`$ and $`F_{x,y,z}^{[n+\frac{1}{2}]}=F(x^{[n+\frac{1}{2}]},y^{[n+\frac{1}{2}]},z^{[n+\frac{1}{2}]},t^{[n+\frac{1}{2}]}).`$
We have tested our integrator for a non-conservative system that includes the Sun and a test particle and is governed by the Poynting-Robertson drag (two values of the parameter $`\beta =`$ 0.285 and 0.057 have been used). As initial conditions for {$`a,e,i`$}, we used the following sets: {39 AU, 0.25, $`10^{}`$} and {45 AU, 0, $`10^{}`$}. The integration time step was taken in the range 16 to 160 days (it was smaller, the closer the test particle to the star). Fig. 7 shows the accuracy of our integrator evaluated by the change of the first dissipative integral of motion with time, $`\delta C=C(t)C(0)`$, where $`C=\frac{a(1e^2)}{e^{4/5}}`$ (see, e.g., GOMT 1997b). It can be seen that, as the particle approaches the star, the integration error increases, but it never exceeds 1% and is much better during the larger part of the trajectory. We emphasize that, since in the real Solar system any dust particle experiences strong gravitational perturbations from the planets so that the particleโs trajectory is highly chaotic, an accumulation of dissipative integral errors as small as shown in Fig. 7 is of no importance.
Thus, our integrator takes into account close approaches with planets, which occur frequently for dust particles spiraling toward the star due to the PR-drag, and demonstrates good results in terms of stability of the integration error. In addition, this integrator is $`1.51.9`$ times faster than the ordinary 2nd-order Runge-Kutta integrator (see Taidakova 1997). Although our integrator is not as speedy as symplectic integrators, it has an important advantage since it is applicable to dissipative systems, along with non-dissipative ones.
We found that our integrator employed in a non-rotating reference frame has a similar accuracy but is several times faster than in a rotating system (Taidakova et al. 2000, in preparation). As an example of the use of our integrator in a non-rotating frame, we have recently simulated a warp observed in the circumstellar disk of Beta Pictoris (Gorkavyi et al. 2000c).
References
Backman, D.E., Dasgupta, A. & Stencel, R.E. Model of a Kuiper belt small grain population and resulting far-infrared emission. 1995, ApJ 450, L35-L38
Brown, W.R. & Luu, J.X. 1998. Properties of model comae around Kuiper belt and Centaur objects. Icarus 135, 415-430
Divine, N. Five populations of interplanetary meteoroids. 1993, J. Geophys. Res. 98E, 17029-17048
Flynn, G.J. 1996, Sources of 10 micron interplanetary dust: the contribution from the Kuiper belt. In Physics, Chemistry, and Dynamics of Interplanetary Dust, ed. B. Gustafson & M. Hanner, (San Francisco: ASP), ASP Conf. Ser. 104, p. 171-175
Fridman, A.M. & Gorkavyi, N.N. 1999, Physics of Planetary Rings. Celestial Mechanics of a Continuous Media. Springer-Verlag, pp. 436
Gorkavyi, N.N., Ozernoy, L.M. & Mather, J.C. 1997a, A new approach to dynamical evolution of interplanetary dust due to gravitational scattering. ApJ 474, 496-502
Gorkavyi, N.N., Ozernoy, L.M., Mather, J.C. & Taidakova, T. 1997b, Quasi-stationary states of dust flows under Poynting-Robertson drag: new analytical and numerical solutions, 1997, ApJ 488, 268-276
Gorkavyi, N.N., Ozernoy, L.M., Mather, J.C. & Taidakova, T. 1998a, Structure of the zodiacal cloud: new analytical and numerical solutions. Earth, Planets and Space, 50, 539-544
Gorkavyi, N.N., Ozernoy, L.M., Mather, J.C. & Taidakova, T. 1998b, The large-scale structures in the Solar system: II. Resonant dust belts associated with the orbits of four giant planets. astro-ph/9812480).
Gorkavyi, N., Ozernoy, L., Mather, J., & Heap, S. 2000a, โOrbital motion of resonant clumps in dusty circumstellar disks as a signature of an embedded planetโ. WWW e-print astro-ph/0005347; In Disks, Planetesimals, and Planetsโ (F. Garzon et al., eds.) ASP Conf. Ser. $`\underset{ยฏ}{\text{000}}`$, 000-000 (in press)
Gorkavyi, N.N., Ozernoy, L.M., Mather, J.C. & Taidakova, T. 2000b, The NGST and the zodiacal light in the Solar system. In NGST Science and Technology Exposition (eds. E.P. Smith & K.S. Long), ASP Series, 207, 462-467 ($``$ astro-ph/9910551)
Gorkavyi, N.N., Heap, S.R., Ozernoy, L.M., Mather, J.C. & Taidakova, T. 2000c, Model for the warp in the $`\beta `$ Pictoris disk (in preparation)
Gurnett, D.A., Anser, J.A. Kurth, W.S., & Granroth, L.J. 1997, Micron-sized particles detected in the outer Solar system by the Voyager 1 and 2 plasma wave instruments. Geophys. Res. Lett. 24, 3125-3128
Gustafson, B.A.S. 1994, Physics of zodiacal light. Ann. Rev. Earth Planet. Sci. 22, 553-595
Hamilton, P.D. 1994, A comparison of Lorentz, planetary gravitational, and satellite gravitational resonances. Icarus, 109, 221-240
Humes, D.H. 1980, Results of Pioneer 10 and 11 meteoroid experiments: interplanetary and near-Saturn. J. Geophys. Res., 85, 5841-5852
Kessler, D.J. 1981, Derivation of the collision probability between orbiting objects: the lifetimes of Jupiterโs outer moons. Icarus 48, 39-48
Kortenkamp, S.J. & Dermott, S.F. 1998, Accretion of interplanetary dust particles by the Earth. Icarus 135, 469-495
Levison, H.F., Duncan M.J. 1997. From the Kuiper belt to Jupiter-family comets: the spatial distribution of ecliptic comets. Icarus 127, 13-32
Liou, J.-C. & Zook, H.A. 1997. Evolution of interplanetary dust particles in mean motion resonances with planets. Icarus, 128, 354-367
Liou, J.-C. & Zook, H.A. 1999, Signatures of the giant planets imprinted on the Edgeworth-Kuiper belt dust disk. Astron. J. 118, 580-590
Liou, J.-C., Zook, H.A., & Dermott, S.F. 1996, Kuiper belt grains as a source of interplanetary dust particles. Icarus 124, 429-440
Liou, J.-C., Zook, H.A. & Jackson, A.A. 1995. Radiation pressure, Poynting-Robertson drag, and solar wind in the restricted three-body problem. Icarus 116, 186-201
Marsden, B.G. 1998, MPEC 1998-v14: Distant Minor Planets
Ozernoy, L.M., Gorkavyi, N.N., Taidakova, T. 2000a, Large scale structures in the outer Solar system: I. Cometary belts with resonant features associated with the giant planets. Mon. Not. R.A.S. (submitted) (OGT 2000a)
Ozernoy, L.M., Gorkavyi, N.N., Taidakova, T., 2000b, Four cometary belts associated with the orbits of giant planets: a new view of the outer Solar systemโs structure emerges from numerical simulations. ACM conference (Cornell Univ., July 26-30, 1999). WWW e-print: astro-ph/0001316; Planetary Space Sci. (in press) (OGT 2000b)
Ozernoy, L.M., Gorkavyi, N.N., Mather, J.C. & Taidakova, T. 2000c, Signatures of Exo-solar Planets in Circumstellar Dust Disks. Astrophys. J. Lett., July 10 issue
Potter, D. 1973, Computational Physics (John Wiley & Sons)
Roques, F., Scholl, H., Sicardy, B. & Smith, B.A. 1994, Is there a planet around $`\beta `$ Pictoris ? Perturbations of a planet on a circumstellar dust disk. 1. The numerical model. Icarus, 108, 37-58
Taidakova, T. 1997, A new stable method for long-time integration in an N-body problem. in Astronomical Data Analyses, Software and Systems VI, ed. G. Hunt & H.E. Payne, (San Francisco: ASP), ASP Conf. Ser. 125, p. 174
Taidakova, T. & Gorkavyi, N.N. 1999. New numerical method for non-conservative systems. The Dynamics of Small Bodies in the Solar Systems: A Major Key to Solar Systems Studies, Eds. B.A. Steves and B.A. Roy, Kluwer Academic Publishers, p. 393
Weidenschilling, S.J. & Jackson, A.A. 1993. Orbital resonances and Poynting-Robertson drag. Icarus 104, 244-254
TABLE 1
Details of Computational Runs
Particle lifetime Number of recorded Number of computed (in Myrs) $`(^1)`$ $`(a,e,i)`$-elements$`(^2)`$ spatial positions (in $`10^6`$) (in $`10^{10}`$) $`\underset{ยฏ}{\text{Grains of }r=5\text{-}10\mu \text{m}}`$ No planets, 20 grains used 10.1 1.2 0.7 Planets included, 80 grains used 20.3 9.9 5.5 $`\underset{ยฏ}{\text{Grains of }r=1\text{-}2\mu \text{m}}`$ No planets, 20 grains used 4.1 0.5 0.3 Planets included, 200 grains used 6.7 8.1 6.0
until the particle impacts the Sun, or a planet, or is ejected from the Solar system. An average value of the lifetime is given. A larger value of the lifetime for larger-size grains is explained by a slower PR-drag. Presence of planets additionally increases the particle lifetime due to captures into resonances.
taken with the time step $`=`$ 1 Neptuneโs revolution about the Sun.
Figure Captions
Figure 1. Representative trajectory of a dust particle of of $`r=5`$-$`10\mu `$m ($`\beta =0.057`$) on the planes of orbital coordinates {$`a,e`$} (panel a) and {$`a,i`$} (panel b). The trajectory presents the particle positions taken every $`5\times 10^3`$ yrs. Dashed curves show the boundaries of the crossing zones of the four giant planets. Diamond indicates the particleโs initial position. Numbers 1 to 8 mark the dominating dynamical process on the given part of the trajectory: 1 โ a โjumpโ from the parent body (KBO) due to the solar pressure (see Eqs. 1 and 2); 2 โ drift of particles due to the P-R drag; 3 โ resonant capture into the 3:2 resonance with Neptune, which results in a balance between the P-R drag and gravitational influence of the planet; 4-8 \- gravitational scattering of the particle by giant planets, with eventual ejection out of the system by Jupiter.
Figure 2. 2D density of the kuiperoidal dust on the plane of orbital coordinates, with bins $`\mathrm{\Delta }a=0.15`$ AU, $`\mathrm{\Delta }e=0.01,\mathrm{\Delta }i=0.5^{}`$. To represent the number of particles in each cell, a decimal-logarithm grey scale is employed, i.e. each shade differs 10-fold from the neighboring one. Numerous resonant lines and gaps are seen. The boundaries of the crossing zones of the four giant planets are indicated by solid curves. Positions of the first 40 (a, b) or 100 (c,d) Kuiper belt objects taken from Marsden (1998) are shown by diamonds. Numerous resonant structures are clearly seen.
a, b. 80 dust particles of $`r=5`$-$`10\mu `$m ($`\beta =0.057`$) start their journey from the pericenters and the apocenters of orbits of 40 kuiperoids.
c, d. 200 dust particles of radius $`r=1`$-$`2\mu `$m ($`\beta =0.285`$) start their journey from the pericenters and the apocenters of orbits of 100 kuiperoids.
Figure 3. Distribution of kuiperoidal dust in semimajor axis, in terms of the number of particles per bin, $`n(a)`$, averaged over each trajectory. The bin size $`\mathrm{\Delta }a=0.01a_{Neptune}=0.3`$ AU. The distributions of small ($`r=1`$-$`2\mu `$m) and large ($`r=5`$-$`10\mu `$m) dust particles are shown by dashed and solid lines, respectively. Various resonant structures are indicated by arrows, which are heavy for large particles and thin for small ones (J, S, and U stand for the Jovian, Saturnian, and Uranian resonances, respectively; all other resonances are with Neptune).
Figure 4. Distribution of kuiperoidal dust in the distance of pericenter, in terms of the number of particles per bin, $`n(a)`$, averaged over each trajectory. The bin size $`\mathrm{\Delta }a=0.3`$ AU. The distributions of small ($`r=1`$-$`2\mu `$m) and large ($`r=5`$-$`10\mu `$m) dust particles are shown by dashed and solid lines, respectively.
Figure 5. Column density of kuiperoidal dust population, in terms of number of particles per 0.3 AU bin averaged over one trajectory, as a function of heliocentric distance (solid line). The distributions of small ($`r=1`$-$`2\mu `$m) and large ($`r=5`$-$`10\mu `$m) dust particles are shown by thin and heavy lines, respectively. Dotted line shows a distribution of large particles obtained with the use of only 25% (i.e. 20) of the available particle trajectories, which indicates that the results depend rather weakly upon the number of trajectories used. Dashed and dashed-dotted lines indicate the surface density of kuiperoidal dust computed from 20 particle trajectories in the absence of planets. Larger-size particles, due to a slower motion, form a denser dust population.
Figure 6. Density profile of a small, $`r=1`$-$`2\mu `$m, kuiperoidal dust perpendicular to the ecliptic plane, $`N(R,Z)`$, at heliocentric distances up to 10 AU. To represent the number of particles in each cell, a natural-logarithm grey scale is used, i.e. each shade differs $`e`$-fold from the neighboring one. A remarkable density minimum between Mars and Jupiter is clearly seen.
Figure 7. An average numerical error in the dissipative integral of motion along the particle trajectory (shown as a function of semimajor axis) for a dust particle drifting under the P-R drag toward the Sun. Trajectories starting from the resonant kuiperoids are marked with 1 and 3, while those starting from the flat component of KBOs are marked with 2 and 4. Initial orbital positions {$`a_0,e_0,i_0`$} of the parent KBOs are as follows:
1a,1p โ {39 AU, 0.25, $`10^{}`$}, start from apocenter and pericenter, respectively; $`\beta `$=0.285.
2 โ {45 AU, 0, $`i=10^{}`$}; $`\beta `$=0.285.
3a,3p โ {39 AU, 0.25, $`i=10^{}`$}, start from apocenter and pericenter, respectively; $`\beta `$=0.057.
4 โ {45 AU, 0, $`i=10^{}`$}; $`\beta `$=0.057.
For an overwhelming majority of trajectories, the integration errors, $`10^310^4`$, are well within the acceptable limits.
|
warning/0006/gr-qc0006076.html
|
ar5iv
|
text
|
# Casimir energy and variational methods in AdS spacetime
## I Introduction
The problem of computing vacuum fluctuations in a field theory can be considered as the first step to probe the validity of a theory. An example is given by the zero point energy (ZPE) responsible of the Casimir effect. This one was predicted by Casimir and experimentally confirmed in the Philips laboratories. This is induced when the presence of electrical conductors distorts the zero point energy of the quantum electrodynamics vacuum. Two parallel conducting surfaces, in a vacuum environment, attract one another by a very weak force that varies inversely as the fourth power of the distance between them. This kind of energy is a pure quantum effect; no real particles are involved, only virtual ones. The difference between the stress-energy computed in presence and in absence of the plates with the same boundary conditions gives
$$\mathrm{\Delta }T^{\mu \nu }=T^{\mu \nu }_{vac}^pT^{\mu \nu }_{vac}=\frac{\pi ^2}{720a^4}\left(\begin{array}{cccc}1& 0& 0& 0\\ 0& 1& 0& 0\\ 0& 0& 1& 0\\ 0& 0& 0& 3\end{array}\right).$$
(1)
It is evident that separately, each contribution coming from the summation over all possible resonance frequencies of the cavities is divergent and devoid of physical meaning but the difference between them in the two situations (with and without the plates) is well defined. Note that the energy density
$$\rho =E/V=\mathrm{\Delta }T^{00}=\frac{\pi ^2}{720a^4}$$
(2)
is negative. One can in general formally define the Casimir energy as follows
$$E_{Casimir}\left[\right]=E_0\left[\right]E_0\left[0\right],$$
(3)
where $`E_0`$ is the zero-point energy and $``$ is a boundary. In General Relativity, at the classical level, there exists a subtraction procedure related to the Arnowitt-Deser-Misner (ADM) approach , namely the ADM energy or mass, which can be improperly thought as the classical aspect of the Casimir energy. In a recent paper, the problem of computing the Casimir energy in presence of the Schwarzschild metric for the gravitational field has been considered. The classical energy associated to the related gravitational Casimir energy is represented by the ADM mass
$$M=\underset{r\mathrm{}}{lim}_\mathrm{\Sigma }\sqrt{\widehat{g}}\widehat{g}^{ij}\left[\widehat{g}_{ik,j}\widehat{g}_{ij,k}\right]๐S^k,$$
(4)
where $`\widehat{g}^{ij}`$ is the metric induced on a spacelike hypersurface $`\mathrm{\Sigma }`$ which has a boundary at infinity like $`S^2`$. An equivalent definition of the classical energy is given by the quasilocal energy defined by
$$E_{q.l.}=\text{ }\frac{1}{8\pi G}_{S^2}d^2x\sqrt{\sigma }\left(kk^0\right),$$
(5)
where $`k`$ is the extrinsic curvature referred to the Schwarzschild space and $`k^0`$ is the extrinsic curvature referred to flat space. $`\sigma `$ is the two-dimensional determinant coming from the induced metric $`\sigma _{ab}`$ on the boundaries $`S^2`$. It is relevant to observe that the Schwarzschild space is asymptotically flat, namely when $`r\mathrm{}`$ we recover the flat metric. In this case to correctly compute the classical energy term a subtraction procedure is involved as widely discussed in Refs.. When we transpose this procedure to one loop calculations, we get the zero point energy (ZPE) for gravitons embedded in flat space
$$2\frac{1}{2}\frac{d^3k}{\left(2\pi \right)^3}\sqrt{k^2}.$$
(6)
This term has a quartic ultra-violet (UV) divergence. The same kind of divergence is present when the Schwarzschild background is considered. However their difference has a divergence degree of a logaritmic type. Since boundary conditions are the same, this ZPEโs difference at one loop represents a Casimir-like computation. In this paper we would like to extend the same evaluation reported in Ref. to the computation of the Casimir-like energy for a Schwarzschild-Anti-de Sitter (S-AdS) space at zero temperature discussing the possible existence of an unstable mode. The reason to compute such a correction comes from an analogy between the Schwarzschild metric and the S-AdS metric. Indeed both metrics are spherically symmetric and are characterized by only one root in the gravitational potential. Moreover, no natural outer boundary is present. The rest of the paper is structured as follows, in section II we define the S-AdS line element, in section III we compute the quasilocal energy and the quasilocal mass for the S-AdS space, in section IV we give some of the basic rules to perform the functional integration and we define the Hamiltonian approximated up to second order, in section V we look for stable modes of the spin-two operator acting on transverse traceless tensors, in section VI we show the existence of only one negative mode, in section VII we find a critical radius below which we have a stabilization of the system. We summarize and conclude in section VIII.
## II The Schwarzschild-Anti-de Sitter metric
The S-AdS line element is defined as
$$ds^2=f\left(r\right)dt^2+f\left(r\right)^1dr^2+r^2d\mathrm{\Omega }^2,$$
(7)
where
$$f\left(r\right)=\left(1\frac{2MG}{r}+\frac{r^2}{b^2}\right),$$
(8)
$`b^2=\sqrt{\frac{3}{\mathrm{\Lambda }_c}}`$ and $`\mathrm{\Lambda }_c<0`$ is the negative cosmological constant. For $`\mathrm{\Lambda }_c=0`$ the metric describes the Schwarzschild metric, while for $`M=0`$, we obtain
$$ds^2=\left(1+\frac{r^2}{b^2}\right)dt^2+\left(1+\frac{r^2}{b^2}\right)^1dr^2+r^2d\mathrm{\Omega }^2,$$
(9)
i.e. the Anti-de Sitter metric (AdS). The gravitational potential $`f\left(r\right)`$ of $`\left(\text{7}\right)`$ has only one root located at
$$\overline{r}=\sqrt[3]{\frac{3MG}{\mathrm{\Lambda }_c}+\sqrt{\frac{1}{\mathrm{\Lambda }_c^2}\left(9\left(MG\right)^2+\frac{1}{\mathrm{\Lambda }_c}\right)}}+\sqrt[3]{\frac{3MG}{\mathrm{\Lambda }_c}\sqrt{\frac{1}{\mathrm{\Lambda }_c^2}\left(9\left(MG\right)^2+\frac{1}{\mathrm{\Lambda }_c}\right)}}$$
(10)
and the gravitational potential can be written as
$$f\left(r\right)=\frac{\left(r\overline{r}\right)\left(r^2+\overline{r}r+\overline{r}^2+b^2\right)}{rb^2}.$$
(11)
From Eq.$`\left(\text{8}\right)`$, evaluated at $`\overline{r}`$, the parameter $`M`$ can be written as
$$MG=\frac{\overline{r}\left(\overline{r}^2+b^2\right)}{2b^2}.$$
(12)
In complete analogy with the Schwarzschild case, we will consider a constant time slice $`\mathrm{\Sigma }`$ of the S-AdS manifold $``$<sup>*</sup><sup>*</sup>*In Appendix A, we will report the details concerning the Kruskal-Szekeres description of the S-AdS manifold.. Even if there is a cosmological constant term we generalize the terminology by saying that the hypersurface $`\mathrm{\Sigma }`$ is an Einstein-Rosen bridge with wormhole topology $`S^2\times R^1`$. The Einstein-Rosen bridge defines a bifurcation surface dividing $`\mathrm{\Sigma }`$ in two parts denoted by $`\mathrm{\Sigma }_+`$ and $`\mathrm{\Sigma }_{}`$. Our purpose is to consider perturbations at $`\mathrm{\Sigma }`$ with $`t`$ constant in absence of matter fields, which naturally define quantum fluctuations of the Einstein-Rosen bridge. The explicit expression of the Hamiltonian can be calculated by means of the following line element
$$ds^2=N^2\left(dx^0\right)^2+g_{ij}\left(N^idx^0+dx^i\right)\left(N^jdx^0+dx^j\right),$$
(13)
where $`N`$ is called the lapse function and $`N_i`$ is the shift function. When $`N=\sqrt{1\frac{2MG}{r}+\frac{r^2}{b^2}}`$, $`N_i=0`$ and
$$g_{ij}dx^idx^j=\left(1\frac{2MG}{r}+\frac{r^2}{b^2}\right)^1dr^2+r^2d\mathrm{\Omega }^2,$$
(14)
we recover the S-AdS line element. On the slice $`\mathrm{\Sigma }`$, deviations from the S-AdS metric spatial section of the form
$$g_{ij}=\overline{g}_{ij}+h_{ij}$$
(15)
will be considered with $`N_i=0`$ and $`NN\left(r\right)`$. Then the line element $`\left(\text{13}\right)`$ becomes
$$ds^2=N^2\left(r\right)\left(dx^0\right)^2+g_{ij}dx^idx^j$$
(16)
and the total Hamiltonian is
$$H_T=H_\mathrm{\Sigma }+H_\mathrm{\Sigma }=_\mathrm{\Sigma }d^3x(N+N_i^i)+H_\mathrm{\Sigma },$$
(17)
where
$$\{\begin{array}{c}=G_{ijkl}\pi ^{ij}\pi ^{kl}\left(\frac{16\pi G}{\sqrt{g}}\right)\left(\frac{\sqrt{g}}{16\pi G}\right)\left(R^{\left(3\right)}+\frac{6}{b^2}\right)\text{ (Super Hamiltonian)}\hfill \\ ^i=2\pi _{|j}^{ij}\text{(Super Momentum)}\hfill \end{array}.$$
(18)
and $`H_\mathrm{\Sigma }`$ represents the energy stored into the boundary. According to Witten, to discuss the existence of an unstable sector, we have to compare spaces with the same boundary conditions. An instability appears when we consider the Euclidean S-AdS spacetime with a periodically identified time representing the equilibrium temperature of a S-AdS black hole with the environment. The same boundary conditions on the AdS spacetime can be imposed if the temperature on the boundary is the same. Indeed the AdS spacetime has no natural temperature and this seems to suggest that only the โhotโ AdS spacetime will be unstable. However, by applying the same method of Ref., it is possible to discuss if the instability appears even when we have the $`T=0`$ temperature case. To this purpose the expression we need to evaluate is
$$E^{SAdS}(M,b)=E^{AdS}\left(b\right)+\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}+\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|1loop}.$$
(19)
$`E^{AdS}\left(b\right)`$ represents the reference space energy which is zero for the flat space. $`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}`$represents the energy difference stored in the boundaries due to the presence of the hole and $`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|1loop}`$ is the quantum correction to a classical term.
## III Quasilocal Energy and Quasilocal Mass for the S-AdS space
In this section we fix our attention on the classical part of Eq.$`\left(\text{19}\right)`$. We consider the outer boundary located at some radius $`R`$. Thus the total energy at the classical level is
$$E^{SAdS}(M,b)=E^{AdS}\left(b\right)+\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}.$$
(20)
We begin by looking at the โoutside regionโ of the Kruskal manifold associated to the S-AdS spacetimeSee Appendix A for details.. We will use the quasilocal energy to evaluate $`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}`$. Quasilocal energy is defined as the value of the Hamiltonian that generates unit time translations orthogonal to the two-dimensional boundary,
$$\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}=\frac{1}{8\pi G}_{S^2}d^2x\sqrt{\sigma }\left(kk^0\right),$$
(21)
where $`\left|N\right|=1`$ at $`S^2`$ and $`k`$ is the trace of the extrinsic curvature corresponding to the S-AdS space and $`k^0`$ is the trace of the extrinsic curvature referred to the AdS space. Following Refs. and by means of Eq.$`\left(\text{8}\right),`$ we obtain
$`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}={\displaystyle \frac{1}{8\pi G}}{\displaystyle _{S^2}}๐\mathrm{\Omega }^2r^2\left[{\displaystyle \frac{2\sqrt{f\left(r\right)}}{r}}+{\displaystyle \frac{2\sqrt{f\left(r\right)_{|M=0}}}{r}}\right]_{|r=R}`$
$$=\frac{R}{G}\left[\sqrt{1\frac{2MG}{r}+\frac{r^2}{b^2}}\sqrt{1+\frac{r^2}{b^2}}\right]\underset{Rb}{}\frac{Mb}{R}.$$
(22)
When the boundary is pushed to infinity $`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}0`$. The same happens to the temperature defined by
$$T=\left(\frac{E}{S}\right)_{r=R}=\frac{1}{2\pi }\frac{\kappa }{N\left(R\right)},$$
(23)
where $`N\left(R\right)=\sqrt{f\left(R\right)}`$ is the redshift factor and $`\kappa `$ is the surface gravity defined by
$$\underset{r\overline{r}}{lim}\frac{1}{2}\left|g_{00}^{^{}}\left(r\right)\right|=\frac{3\overline{r}^2+b^2}{2\overline{r}b^2}.$$
(24)
Thus except the limiting case of pushing the boundary to infinity, the temperature $`T`$ and the classical energy $`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}`$ do not vanish and therefore we cannot consider the problem of searching for unstable modes at zero temperature. Nevertheless if we look at the whole S-AdS manifold, the total classical energy can be written as
$`E^{SAdS}(M,b)=E^{AdS}\left(b\right)+E_{tot}(M,b)`$
$$=E^{AdS}\left(b\right)+\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}^++\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}^{}$$
(25)
with
$`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}^+={\displaystyle \frac{1}{8\pi G}}{\displaystyle _{S_+^2}}d^2x\sqrt{\sigma }\left(kk^0\right),`$
$$\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}^{}=\frac{1}{8\pi G}_{S_{}^2}d^2x\sqrt{\sigma }\left(kk^0\right),$$
(26)
and $`\left|N\right|=1`$ at both $`S_+^2`$ and $`S_{}^2`$. $`E_{tot}(M,b)`$ is the quasilocal energy of a spacelike hypersurface $`\mathrm{\Sigma }=\mathrm{\Sigma }_+\mathrm{\Sigma }_{}`$ bounded by two boundaries $`S_+^2`$ and $`S_{}^2`$ located in the two disconnected regions $`_+`$ and $`_{}`$ respectively. To evaluate $`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|classical}^\pm `$ we can use Eq.$`\left(\text{22}\right)`$ or more pictorially by looking at the static Einstein-Rosen bridge associated to the S-AdS space, whose metric is
$$ds^2=N^2\left(r\right)dt^2+g_{xx}dx^2+r^2\left(x\right)d\mathrm{\Omega }^2,$$
(27)
where $`N`$, $`g_{xx}`$, and $`r`$ are functions of the radial coordinate $`x`$ continuously defined on $``$, with
$$dx=\pm \frac{dr}{\sqrt{1\frac{2MG}{r}+\frac{r^2}{b^2}}},$$
(28)
where the plus sign is relative to $`\mathrm{\Sigma }_+`$, while the minus sign is related to $`\mathrm{\Sigma }_{}`$. If we make the identification $`N^2=1\frac{2MG}{r}+\frac{r^2}{b^2}`$, the line element $`\left(\text{27}\right)`$ reduces to the S-AdS metric written in another form. The boundaries $`S_\pm ^2`$ are located at coordinate values $`x=\overline{x}^\pm `$. The normal to the boundaries is $`n^\mu =\left(h^{xx}\right)^{\frac{1}{2}}\delta _y^\mu `$. By using the expression of the trace
$$k=\frac{1}{\sqrt{h}}\left(\sqrt{h}n^\mu \right)_{,\mu },$$
(29)
we obtain
$$k^{SAdS}=\{\begin{array}{c}2r,_x/ron\mathrm{\Sigma }_+\\ 2r,_x/ron\mathrm{\Sigma }_{}\end{array}.$$
(30)
Thus the computation of $`E_+`$ gives exactly the result of Eq.$`\left(\text{22}\right)`$. On the other hand the computation of $`E_{}`$ gives the same value but with the reversed sign. Thus one getsNote that if we take as a reference space the flat space, then the trace is taken to be $`k^{flat}=2/r`$ and
$`\left(E^{SAdS}E^{flat}\right)_\pm =\{\begin{array}{c}R^2/GbS_+^2\\ R^2/GbS_{}^2\end{array}.`$
When $`R\mathrm{}`$, $`\left(E^{SAdS}E^{flat}\right)_\pm \mathrm{}`$.
$$\left(E^{SAdS}E^{AdS}\right)_\pm =\{\begin{array}{c}Mb/RonS_+^2\\ Mb/RonS_{}^2\end{array},$$
(31)
where for $`E_{}`$ we have used the conventions relative to $`\mathrm{\Sigma }_{}`$ and $`S_{}^2`$. Therefore for every value of the boundary $`R`$, (provided we take symmetric boundary conditions with respect to the bifurcation surface, even for the limiting value $`R\mathrm{}`$), we have
$$E^{SAdS}(M,b)=E^{AdS}\left(b\right)+Mb/RMb/R=E^{AdS}\left(b\right),$$
(32)
namely the energy is conserved for every choice of the boundary location.
## IV Energy Density Calculation in Schrรถdinger Representation
In previous section we have fixed our attention to the classical part of Eq.$`\left(\text{19}\right)`$. In this section, we apply the same calculation scheme of Refs. to compute one loop corrections to the classical S-AdS term. Like the Schwarzschild case, there appear two classical constraints for the Hamiltonian
$$\{\begin{array}{c}\text{ }=0\hfill \\ ^i=0\hfill \end{array},$$
(33)
which are satisfied both by the S-AdS and AdS metric and two quantum constraints
$$\{\begin{array}{c}\stackrel{~}{\mathrm{\Psi }}\text{ }=0\hfill \\ ^i\stackrel{~}{\mathrm{\Psi }}=0\hfill \end{array}.$$
(34)
$`\stackrel{~}{\mathrm{\Psi }}`$ $`=0`$ is known as the Wheeler-DeWitt equation (WDW). Nevertheless, our purpose is the computation of
$$\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|1loop}=\frac{\mathrm{\Psi }\left|H_\mathrm{\Sigma }^{SAdS}H_\mathrm{\Sigma }^{AdS}\right|\mathrm{\Psi }}{\mathrm{\Psi }|\mathrm{\Psi }}$$
(35)
where $`H_\mathrm{\Sigma }^{SAdS}`$ and $`H_\mathrm{\Sigma }^{AdS}`$ are the total Hamiltonians referred to the S-AdS and AdS spacetimes respectively for the volume term and $`\mathrm{\Psi }`$ is a wave functional obtained following the usual WKB expansion of the WDW solution. In this context, the approximated wave functional will be substituted by a trial wave functional of the gaussian form according to the variational approach we shall use to evaluate $`\mathrm{\Delta }E_{AdS}^{SAdS}(M,b)_{|1loop}`$. Following the same procedure of Refs., we expand the three-scalar curvature $`d^3x\sqrt{g}\left(R^{\left(3\right)}+6/b^2\right)`$ up to $`o\left(h^2\right)`$ and we get
$`{\displaystyle _\mathrm{\Sigma }}d^3x\sqrt{\overline{g}}\left[{\displaystyle \frac{1}{4}}h\mathrm{}h+{\displaystyle \frac{1}{4}}h^{li}\mathrm{}h_{li}{\displaystyle \frac{1}{2}}h^{ij}_l_ih_j^l+{\displaystyle \frac{1}{2}}h_l_ih^{li}{\displaystyle \frac{1}{2}}h^{ij}R_{ia}h_j^a+{\displaystyle \frac{1}{2}}hR_{ij}h^{ij}\right]`$
$$_\mathrm{\Sigma }d^3x\sqrt{\overline{g}}\left[\frac{1}{4}h^{li}\left(6/b^2\right)h_{li}\right].$$
(36)
To explicitly make calculations, we need an orthogonal decomposition for both $`\pi _{ij\text{ }}`$and $`h_{ij}`$ to disentangle gauge modes from physical deformations. We define the inner product
$$h,k:=_\mathrm{\Sigma }\sqrt{g}G^{ijkl}h_{ij}\left(x\right)k_{kl}\left(x\right)d^3x,$$
(37)
by means of the inverse WDW metric $`G_{ijkl}`$, to have a metric on the space of deformations, i.e. a quadratic form on the tangent space at h, with
$$\begin{array}{c}G^{ijkl}=(g^{ik}g^{jl}+g^{il}g^{jk}2g^{ij}g^{kl})\text{.}\end{array}$$
(38)
The inverse metric is defined on co-tangent space and it assumes the form
$$p,q:=_\mathrm{\Sigma }\sqrt{g}G_{ijkl}p^{ij}\left(x\right)q^{kl}\left(x\right)d^3x\text{,}$$
(39)
so that
$$G^{ijnm}G_{nmkl}=\frac{1}{2}\left(\delta _k^i\delta _l^j+\delta _l^i\delta _k^j\right).$$
(40)
Note that in this scheme the โinverse metricโ is actually the WDW metric defined on phase space. The desired decomposition on the tangent space of 3-metric deformations is:
$$h_{ij}=\frac{1}{3}hg_{ij}+\left(L\xi \right)_{ij}+h_{ij}^{}$$
(41)
where the operator $`L`$ maps $`\xi _i`$ into symmetric tracefree tensors
$$\left(L\xi \right)_{ij}=_i\xi _j+_j\xi _i\frac{2}{3}g_{ij}\left(\xi \right).$$
(42)
Thus the inner product between three-geometries becomes
$`h,h:={\displaystyle _\mathrm{\Sigma }}\sqrt{g}G^{ijkl}h_{ij}\left(x\right)h_{kl}\left(x\right)d^3x=`$
$$_\mathrm{\Sigma }\sqrt{g}\left[\frac{2}{3}h^2+\left(L\xi \right)^{ij}\left(L\xi \right)_{ij}+h^{ij}h_{ij}^{}\right].$$
(43)
With the orthogonal decomposition in hand we can define the trial wave functional
$$\mathrm{\Psi }\left[h_{ij}\left(\stackrel{}{x}\right)\right]=๐ฉ\mathrm{exp}\left\{\frac{1}{4l_p^2}\left[hK^1h_{x,y}^{}+\left(L\xi \right)K^1\left(L\xi \right)_{x,y}^{}+hK^1h_{x,y}^{Trace}\right]\right\},$$
(44)
where $`๐ฉ`$ is a normalization factor. Since we are interested only to the perturbations of the physical degrees of freedom, we will fix our attention only to the TT tensor sector reducing therefore the previous form into
$$\mathrm{\Psi }\left[h_{ij}\left(\stackrel{}{x}\right)\right]=๐ฉ\mathrm{exp}\left\{\frac{1}{4l_p^2}hK^1h_{x,y}^{}\right\}.$$
(45)
This restriction is motivated by the fact that if an instability appears this will be in the physical sector referred to TT tensors, namely a non conformal instability. This means that does not exist a gauge choice that can eliminate negative modes. This choice seems also corroborated by the action decomposition of Ref., where only TT tensors contribute to the partition function<sup>ยง</sup><sup>ยง</sup>ยงSee also Ref. for another point of view.. Therefore to calculate the energy density, we need to know the action of some basic operators on $`\mathrm{\Psi }\left[h_{ij}\right]`$. The action of the operator $`h_{ij}`$ on $`|\mathrm{\Psi }=\mathrm{\Psi }\left[h_{ij}\right]`$ is realized by
$$h_{ij}\left(x\right)|\mathrm{\Psi }=h_{ij}\left(\stackrel{}{x}\right)\mathrm{\Psi }\left[h_{ij}\right].$$
(46)
The action of the operator $`\pi _{ij}`$ on $`|\mathrm{\Psi }`$, in general, is
$$\pi _{ij}\left(x\right)|\mathrm{\Psi }=i\frac{\delta }{\delta h_{ij}\left(\stackrel{}{x}\right)}\mathrm{\Psi }\left[h_{ij}\right].$$
(47)
The inner product is defined by the functional integration:
$$\mathrm{\Psi }_1\mathrm{\Psi }_2=\left[๐h_{ij}\right]\mathrm{\Psi }_1^{}\left\{h_{ij}\right\}\mathrm{\Psi }_2\left\{h_{kl}\right\},$$
(48)
and by applying previous functional integration rules, we obtain the expression of the one-loop-like Hamiltonian form for TT (traceless and transverseless) deformations
$$H^{}=\frac{1}{4l_p^2}_{}d^3x\sqrt{g}G^{ijkl}\left[K^1(x,x)_{ijkl}+\left(\mathrm{}_2\right)_j^aK^{}(x,x)_{iakl}\right].$$
(49)
The propagator $`K^{}(x,x)_{iakl}`$ comes from a functional integration and it can be represented as
$$K^{}(\stackrel{}{x},\stackrel{}{y})_{iakl}:=\underset{N}{}\frac{h_{ia}^{}\left(\stackrel{}{x}\right)h_{kl}^{}\left(\stackrel{}{y}\right)}{2\lambda _N\left(p\right)},$$
(50)
where $`h_{ia}^{}\left(\stackrel{}{x}\right)`$ are the eigenfunctions of $`\mathrm{}_{2j}^a`$ and $`\lambda _N\left(p\right)`$ are infinite variational parameters.
## V The Schwarzschild-Anti-de Sitter Metric spin 2 operator and the evaluation of the energy density
The Spin-two operator for the S-AdS metric is defined by
$$\left(\mathrm{}_2\right)_j^a:=\mathrm{}\delta _j^a+2R_j^a+6/b^2\delta _j^a$$
(51)
where $`\mathrm{}`$ is the curved Laplacian (Laplace-Beltrami operator) on a S-AdS background and $`R_{j\text{ }}^a`$ is the mixed Ricci tensor whose components are:
$$R_i^a=\{\frac{2MG}{r^3}2/b^2,\frac{MG}{r^3}2/b^2,\frac{MG}{r^3}2/b^2\}.$$
(52)
Note that the form of the mixed Ricci tensor for the S-AdS space is the same of the mixed Ricci tensor computed in the Schwarzschild space, except for the presence of the negative cosmological term. We are led to study the following eigenvalue equation
$$\left(\mathrm{}\delta _j^a+2R_j^a+6/b^2\delta _j^a\right)h_a^i=E^2h_j^i$$
(53)
where $`E^2`$ is the eigenvalue of the corresponding equation. In doing so, we follow Regge and Wheeler in analyzing the equation as modes of definite frequency, angular momentum and parity. The quantum number corresponding to the projection of the angular momentum on the z-axis will be set to zero. This choice will not alter the contribution to the total energy since we are dealing with a spherical symmetric problem. In this case, Regge-Wheeler decomposition shows that the even-parity three-dimensional perturbation is
$$h_{ij}^{even}(r,\vartheta ,\varphi )=diag[H\left(r\right)\left(1\frac{2MG}{r}+\frac{r^2}{b^2}\right)^1,r^2K\left(r\right),r^2\mathrm{sin}^2\vartheta K\left(r\right)]Y_{l0}(\vartheta ,\varphi ).$$
(54)
Representation $`\left(\text{54}\right)`$ shows a gravitational perturbation decoupling. For a generic value of the angular momentum $`L`$, one gets
$$\{\begin{array}{c}\left(\mathrm{}_l\frac{4MG}{r^3}+\frac{2}{b^2}\right)H\left(r\right)=E_l^2H\left(r\right)\\ \\ \left(\mathrm{}_l+\frac{2MG}{r^3}+\frac{2}{b^2}\right)K\left(r\right)=E_l^2K\left(r\right).\end{array}$$
(55)
The Laplacian restricted to $`\mathrm{\Sigma }`$ can be written as
$$\mathrm{}_l=\left(1\frac{2MG}{r}+\frac{r^2}{b^2}\right)\frac{d^2}{dr^2}+\left(\frac{2r3MG}{r^2}+3\frac{r}{b^2}\right)\frac{d}{dr}\frac{l\left(l+1\right)}{r^2}.$$
(56)
Defining reduced fields
$$H\left(r\right)=\frac{h\left(r\right)}{r};K\left(r\right)=\frac{k\left(r\right)}{r},$$
(57)
and passing to the proper geodesic distance from the throat of the bridge defined by Eq.$`\left(\text{28}\right)`$, the system $`\left(\text{55}\right)`$ becomesThe system does not change in form if we make the minus choice in Eq.$`\left(\text{28}\right)`$.
$$\{\begin{array}{c}\frac{d^2}{dx^2}h\left(x\right)+\left(V^{}\left(x\right)+\frac{3}{b^2}\right)h\left(x\right)=E_l^2h\left(x\right)\\ \\ \frac{d^2}{dx^2}k\left(x\right)+\left(V^+\left(x\right)+\frac{3}{b^2}\right)k\left(x\right)=E_l^2k\left(x\right)\end{array}$$
(58)
with
$$V^{}\left(x\right)=\frac{l\left(l+1\right)}{r^2\left(x\right)}\frac{3MG}{r\left(x\right)^3}.$$
(59)
When $`r\mathrm{}`$, $`x\left(r\right)b\mathrm{ln}r`$ and $`V\left(x\right)0`$. When $`rr_0`$, $`x\left(r\right)0`$ and
$$V^{}\left(x\right)\frac{l\left(l+1\right)}{r_0^2}\frac{3MG}{r_0^3}=const,$$
(60)
where $`r_0`$ satisfies the condition $`r_0>\overline{r}`$. The solution of $`\left(\text{58}\right)`$, in both cases (S-AdS and AdS one) is
$$h\left(px\right)=k\left(px\right)=\sqrt{\frac{2}{\pi }}\mathrm{sin}\left(px\right).$$
(61)
This choice is dictated by the requirement that
$$h\left(x\right),k\left(x\right)0\text{ when }x0\left(\text{alternatively }r\overline{r}\right).$$
(62)
Thus the propagator becomes
$$K_\pm ^{}(x,y)=\frac{V}{2\pi ^2}_0^{\mathrm{}}๐pp^2\frac{\mathrm{sin}\left(px\right)}{r\left(x\right)}\frac{\mathrm{sin}\left(py\right)}{r\left(y\right)}\frac{Y_{l0}(\vartheta ,\varphi )Y_{l^{}0}(\vartheta ,\varphi )}{\lambda _\pm \left(p\right)}$$
(63)
$`\lambda _\pm \left(p\right)`$ is referred to the potential function $`V^\pm \left(x\right)`$. Substituting Eq.$`\left(\text{63}\right)`$ in Eq.$`\left(\text{49}\right)`$ one gets (after normalization in spin space and after a rescaling of the fields in such a way as to absorb $`l_p^2`$)
$$E(M,b,\lambda )=\frac{V}{8\pi ^2}\underset{l=0}{\overset{\mathrm{}}{}}\underset{i=1}{\overset{2}{}}_0^{\mathrm{}}๐pp^2\left[\lambda _i\left(p\right)+\frac{E_i^2(p,M,b,l)}{\lambda _i\left(p\right)}\right]$$
(64)
where
$$E_{1,2}^2(p,M,b,l)=p^2+\frac{l\left(l+1\right)}{r_0^2}\frac{3MG}{r_0^3}+\frac{3}{b^2},$$
(65)
$`\lambda _i\left(p\right)`$ are variational parameters corresponding to the eigenvalues for a (graviton) spin-two particle in an external field and $`V`$ is the volume of the system. By minimizing $`\left(\text{64}\right)`$ with respect to $`\lambda _i\left(p\right)`$ one obtains $`\overline{\lambda }_i\left(p\right)=\left[E_i^2(p,M,b,l)\right]^{\frac{1}{2}}`$ and
$$E(M,b,\overline{\lambda })=\frac{V}{8\pi ^2}\underset{l=0}{\overset{\mathrm{}}{}}\underset{i=1}{\overset{2}{}}_0^{\mathrm{}}๐pp^22\sqrt{E_i^2(p,M,b,l)}\text{ }$$
(66)
with
$`p^2+{\displaystyle \frac{l\left(l+1\right)}{r_0^2}}+{\displaystyle \frac{3}{b^2}}>{\displaystyle \frac{3MG}{r_0^3}}.`$
For the S-AdS background we get
$$E(M,b)=\frac{V}{4\pi ^2}\underset{l=0}{\overset{\mathrm{}}{}}_0^{\mathrm{}}๐pp^2\left(\sqrt{p^2+c_{}^2}+\sqrt{p^2+c_+^2}\right)$$
(67)
where
$`c_{}^2={\displaystyle \frac{l\left(l+1\right)}{r_0^2}}{\displaystyle \frac{3MG}{r_0^3}}+{\displaystyle \frac{3}{b^2}},`$
while when we refer to the AdS space we put $`M=0`$ and $`c^2=`$ $`\frac{l\left(l+1\right)}{r_0^2}+\frac{3}{b^2}`$. Then
$$E\left(b\right)=\frac{V}{4\pi ^2}\underset{l=0}{\overset{\mathrm{}}{}}_0^{\mathrm{}}๐pp^2\left(2\sqrt{p^2+c^2}\right)$$
(68)
Now, we are in position to compute the difference between $`\left(\text{67}\right)`$ and $`\left(\text{68}\right)`$. Since we are interested in the $`UV`$ limit, we have
$`\mathrm{\Delta }E(M,b)=E(M,b)E\left(b^2\right)`$
$`={\displaystyle \frac{V}{4\pi ^2}}{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle _0^{\mathrm{}}}๐pp^2\left[\sqrt{p^2+c_{}^2}+\sqrt{p^2+c_+^2}2\sqrt{p^2+c^2}\right]`$
$$=\frac{V}{4\pi ^2}\underset{l=0}{\overset{\mathrm{}}{}}_0^{\mathrm{}}๐pp^3\left[\sqrt{1+\left(\frac{c_{}}{p}\right)^2}+\sqrt{1+\left(\frac{c_+}{p}\right)^2}2\sqrt{1+\left(\frac{c}{p}\right)^2}\right]$$
(69)
and for $`p^2>>c_{}^2,c^2`$, we obtain
$`{\displaystyle \frac{V}{4\pi ^2}}{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle _0^{\mathrm{}}}dpp^3[1+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{c_{}}{p}}\right)^2{\displaystyle \frac{1}{8}}\left({\displaystyle \frac{c_{}}{p}}\right)^4+1+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{c_+}{p}}\right)^2{\displaystyle \frac{1}{8}}\left({\displaystyle \frac{c_+}{p}}\right)^4`$
$$2\left(\frac{c}{p}\right)^2+\frac{1}{4}\left(\frac{c}{p}\right)^4]=\frac{V}{2\pi ^2}\frac{c_M^4}{8}_0^{\mathrm{}}\frac{dp}{p},$$
(70)
where $`c_M^2=3MG/r_0^3`$. We will use a cut-off $`\mathrm{\Lambda }`$ to keep under control the $`UV`$ divergence
$$_0^{\mathrm{}}\frac{dp}{p}_0^{\frac{\mathrm{\Lambda }}{c_M}}\frac{dx}{x}\mathrm{ln}\left(\frac{\mathrm{\Lambda }}{c_M}\right),$$
(71)
where $`\mathrm{\Lambda }m_p.`$ Thus $`\mathrm{\Delta }E(M,b)`$ for high momenta becomes
$$\mathrm{\Delta }E(M,b)\frac{V}{2\pi ^2}\frac{c_M^4}{16}\mathrm{ln}\left(\frac{\mathrm{\Lambda }^2}{c_M^2}\right)=\frac{V}{32\pi ^2}\left(\frac{3MG}{r_0^3}\right)^2\mathrm{ln}\left(\frac{r_0^3\mathrm{\Lambda }^2}{3MG}\right).$$
(72)
and Eq.$`\left(\text{19}\right)`$ to one loop is
$$E^{SAdS}(M,b)E^{AdS}\left(b\right)=\frac{V}{32\pi ^2}\left(\frac{3MG}{r_0^3}\right)^2\mathrm{ln}\left(\frac{r_0^3\mathrm{\Lambda }^2}{3MG}\right).$$
(73)
Like the Schwarzschild case, we observe that
$$\underset{M0}{lim}\underset{r\overline{r}}{lim}\mathrm{\Delta }E(M,b)\underset{r\overline{r}}{lim}\underset{M0}{lim}\mathrm{\Delta }E(M,b).$$
(74)
This behavior seems to confirm that quantum effects come into play when we try to reach the horizon. By means of Eq.$`\left(\text{12}\right)`$, $`\mathrm{\Delta }E(M,b)`$ becomes
$$\mathrm{\Delta }E(\overline{r},b)=\frac{V}{32\pi ^2}\left(3\frac{\overline{r}\left(\overline{r}^2+b^2\right)}{2b^2r_0^3}\right)^2\mathrm{ln}\left(\frac{2b^2r_0^3\mathrm{\Lambda }^2}{3\overline{r}\left(\overline{r}^2+b^2\right)}\right).$$
(75)
If we set $`\overline{r}=b/\sqrt{3}=r_m`$, which is the location of the minimum of the surface gravity, we find that $`\mathrm{\Delta }E(\overline{r},b)`$ is reduced to
$$\mathrm{\Delta }E\left(b\right)=\frac{V}{32\pi ^2}\left(\frac{2}{\sqrt{3}r_0^3}b\right)^2\mathrm{ln}\left(\frac{\sqrt{3}r_0^3\mathrm{\Lambda }^2}{2b}\right).$$
(76)
Note that in the terminology of the black hole thermodynamics $`r_m`$ corresponds to the unique black hole solution whose temperature reaches its minimum. To better appreciate the result obtained in Eq.$`\left(\text{73}\right)`$, we define a scale variable $`x=3MG/\left(r_0^3\mathrm{\Lambda }^2\right)`$ in such a way that $`\mathrm{\Delta }E\left(M\right)`$ can be cast in the form
$$\mathrm{\Delta }E\left(x\right)=\frac{V}{32\pi ^2}\mathrm{\Lambda }^4x^2\mathrm{ln}x.$$
(77)
A stationary point is reached for $`x=0`$, namely the AdS space and another stationary point is in $`x=e^{\frac{1}{2}}.`$ This last one represents a minimum of $`\mathrm{\Delta }E\left(x\right)`$. This means that there is a probability that the spacetime without the hole will decay into a spacetime with a hole (not black hole). To see if this is really possible, we have to establish if there exist unstable modes. However in case of Eq.$`\left(\text{76}\right)`$ the claim that the stationary point $`x=0`$ represents the AdS space is more delicate. Indeed this corresponds to the vanishing of the parameter $`b`$, leading to a diverging negative cosmological constant, saturating the whole space.
## VI Searching for negative modes
In this paragraph we look for negative modes of the eigenvalue equation $`\left(\text{51}\right)`$. For this purpose we restrict the analysis to the S wave. Indeed, in this state the centrifugal term is absent and this gives the function $`V\left(x\right)`$ a potential well form, which is different when the angular momentum $`l1`$. Moreover the potential well appears only for the $`H`$ component, whose eigenvalue equation is
$$\left(\mathrm{}\frac{4MG}{r^3}+\frac{2}{b^2}\right)H\left(r\right)=E^2H\left(r\right).$$
(78)
$`\mathrm{}`$ is the operator $`\mathrm{}_l`$ of Eq.$`\left(\text{56}\right)`$ with $`l=0`$ and $`E^2>0`$. By defining the reduced field $`h\left(r\right)=H\left(r\right)r`$, Eq.$`\left(\text{78}\right)`$ becomes
$$\frac{d}{dr}\left(\sqrt{1\frac{2MG}{r}+\frac{r^2}{b^2}}\frac{dh}{dr}\right)+\left(\frac{3MG}{r^3}+\stackrel{~}{E}^2\right)\frac{h}{\sqrt{1\frac{2MG}{r}+\frac{r^2}{b^2}}}=0,$$
(79)
where $`\stackrel{~}{E}^2=3/b^2+E^2`$. By means of Eq.$`\left(\text{28}\right)`$, one gets
$`{\displaystyle \frac{dx}{dr}}{\displaystyle \frac{d}{dx}}\left(\sqrt{1{\displaystyle \frac{2MG}{r}}+{\displaystyle \frac{r^2}{b^2}}}{\displaystyle \frac{dh}{dx}}{\displaystyle \frac{dx}{dr}}\right)+\left({\displaystyle \frac{3MG}{r^3}}+\stackrel{~}{E}^2\right){\displaystyle \frac{h}{\sqrt{1\frac{2MG}{r}+\frac{r^2}{b^2}}}}`$
$$=\frac{d}{dx}\left(\frac{dh}{dx}\right)+\left(\frac{3MG}{r^3}+\stackrel{~}{E}^2\right)h=0.$$
(80)
Near the throat
$$x\left(r\right)\frac{\sqrt{2\overline{r}}}{\sqrt{\kappa }}\sqrt{\left(\frac{r}{\overline{r}}1\right)},$$
(81)
where $`\kappa `$ is the surface gravity. By defining the dimensionless variable $`\rho =\frac{r}{\overline{r}}`$, we obtain $`\rho =1+y^2`$ where
$$y=\sqrt{\kappa }x/\sqrt{2\overline{r}}=\stackrel{~}{\kappa }x.$$
(82)
Then Eq.$`\left(\text{80}\right)`$ becomes
$`{\displaystyle \frac{d}{dy}}\left({\displaystyle \frac{dh}{dy}}\right)\stackrel{~}{\kappa }^2+\left({\displaystyle \frac{3MG}{\left(\overline{r}\right)^3\rho ^3\left(y\right)}}+\stackrel{~}{E}^2\right)h`$
$$=\frac{d^2h}{dy^2}+\left(\frac{3MG}{\stackrel{~}{\kappa }^2\left(\overline{r}\right)^3\left(1+y^2\right)^3}+\lambda \right)h=0,$$
(83)
where $`\lambda =\stackrel{~}{E}^2/\stackrel{~}{\kappa }^2`$. Expanding the potential around $`y=0`$, one gets
$$\frac{d^2h}{dy^2}+\left(\frac{3MG}{\stackrel{~}{\kappa }^2\left(\overline{r}\right)^3}\left(13y^2\right)+\lambda \right)h$$
(84)
$$=\frac{d^2h}{dy^2}+\left(\omega ^2y^2\frac{3MG}{\stackrel{~}{\kappa }^2\left(\overline{r}\right)^3}+\lambda \right)h=0,$$
(85)
where $`\omega =\sqrt{9MG/\left(\stackrel{~}{\kappa }^2\left(\overline{r}\right)^3\right)}`$. In this approximation we have obtained the equation of a quantum harmonic oscillator equation whose spectrum is $`E_n=\mathrm{}\omega \left(n+\frac{1}{2}\right)`$. Since we are using natural units, we set $`\mathrm{}=1`$ and
$$\lambda _n=3MG/\left(\stackrel{~}{\kappa }^2\left(\overline{r}\right)^3\right)\sqrt{9MG/\left(\stackrel{~}{\kappa }^2\left(\overline{r}\right)^3\right)}\left(n+\frac{1}{2}\right).$$
(86)
After some algebraic calculation, we obtain
$$\lambda _n=6\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}\left(\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}\frac{1}{\sqrt{2}}\left(n+\frac{1}{2}\right)\right),$$
(87)
where we have used the relation $`\left(\text{12}\right)`$. We see that
$$\lambda _0=6\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}\left(\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}\frac{1}{2\sqrt{2}}\right).$$
(88)
Since the eigenvalue must be positive, the following inequality must hold
$$\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}>\frac{1}{2\sqrt{2}}7b^2+5\overline{r}^2>0,$$
(89)
which is always verified. To proof that there is only one eigenvalue, we look at the second eigenvalue
$$\lambda _1=6\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}\left(\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}\frac{3}{2\sqrt{2}}\right).$$
(90)
The inequality
$$\sqrt{\frac{b^2+\overline{r}^2}{b^2+3\overline{r}^2}}>\frac{3}{2\sqrt{2}}b^2+19\overline{r}^2<0,$$
(91)
which is never verified since $`b`$ and $`\overline{r}`$ are real quantities. Thus we can conclude that there is only one eigenvalue and according to Coleman , this is a signal of a transition from a false vacuum to a true one. The same unstable mode appears also when we introduce a temperature and we look at the thermodynamic stability of a S-AdS black hole within isothermal cavities. In terms of $`E^2`$, the eigenvalue is
$`E^2=3/b^23{\displaystyle \frac{b^2+\overline{r}^2}{2b^2\overline{r}^2}}+{\displaystyle \frac{3}{4b^2\overline{r}^2}}\sqrt{\left(b^2+\overline{r}^2\right)\left(b^2+3\overline{r}^2\right)}`$
$$=3\frac{b^2+3\overline{r}^2}{2b^2\overline{r}^2}+\frac{3}{4b^2\overline{r}^2}\sqrt{\left(b^2+\overline{r}^2\right)\left(b^2+3\overline{r}^2\right)}.$$
(92)
## VII Boundary Reduction and stability
An equivalent approach to Eq.$`\left(\text{79}\right)`$ can be set up by means of a variational procedure applied on a functional whose minimum represents the solution of the problem. Let us define
$`J(h,E^2)={\displaystyle \frac{1}{2}}{\displaystyle \underset{0}{\overset{x\left(a\right)}{}}}๐x\left[\left({\displaystyle \frac{dh\left(x\right)}{dx}}\right)^2{\displaystyle \frac{3MG}{r^3\left(x\right)}}h^2\left(x\right)\right]+{\displaystyle \frac{\stackrel{~}{E}^2}{2}}{\displaystyle \underset{0}{\overset{x\left(a\right)}{}}}๐xh^2\left(x\right),`$
where $`dx`$ is given by Eq.$`\left(\text{28}\right)`$. Eq.$`\left(\text{79}\right)`$ is equivalent to finding the minimum of
$$\stackrel{~}{E}^2=\frac{\underset{0}{\overset{x\left(a\right)}{}}๐x\left[\left(\frac{dh\left(x\right)}{dx}\right)^2\frac{3MG}{r^3\left(x\right)}h^2\left(x\right)\right]}{\underset{0}{\overset{x\left(a\right)}{}}๐xh^2\left(x\right)}.$$
(93)
For future purposes, we use the boundary conditions
$$h\left(x\left(a\right)\right)=0.$$
(94)
When $`r\mathrm{}x\mathrm{}`$ and Eq.$`\left(\text{80}\right)`$ becomes
$$\frac{d^2h}{dx^2}+\stackrel{~}{E}^2h=0$$
(95)
whose asymptotic solution isAlthough the asymptotic behaviour is such that
$`x\left(r\right)=\pm b\mathrm{ln}r,`$
for practical purposes, we prefer to use the variable $`x\left(r\right)`$.
$$h\left(x\right)=A\mathrm{exp}\left(\stackrel{~}{E}x\right)+B\mathrm{exp}\left(\stackrel{~}{E}x\right).$$
(96)
Since asymptotically $`\mathrm{exp}\left(\stackrel{~}{E}x\right)`$ diverges, we set $`B=0`$, then $`h\left(x\right)`$ becomes $`A\mathrm{exp}\left(\stackrel{~}{E}x\right)`$. If we change the variables in a dimensionless form like Eq.$`\left(\text{82}\right)`$, we get
$$\mu =\frac{\stackrel{~}{E}^2}{\stackrel{~}{\kappa }^2}=\frac{\underset{0}{\overset{\overline{y}}{}}๐y\left[\left(\frac{dh\left(y\right)}{dy}\right)^2\frac{3MG}{\overline{r}^3\stackrel{~}{\kappa }^2\rho ^3\left(y\right)}h^2\left(y\right)\right]}{\underset{0}{\overset{y\left(a\right)}{}}๐yh^2\left(y\right)}.$$
(97)
The asymptotic behaviour of $`h\left(x\right)`$ suggests to choose $`h(\lambda ,y)=\mathrm{exp}\left(\lambda y\right)`$ as a trial function, and Eq.$`\left(\text{97}\right)`$ becomes
$$\mu \left(\lambda \right)=\lambda ^2\frac{3MG}{\overline{r}^3\stackrel{~}{\kappa }^2}\frac{\underset{0}{\overset{\overline{y}}{}}\frac{dy}{\rho ^3\left(y\right)}\mathrm{exp}\left(2\lambda y\right)}{\frac{1\mathrm{exp}\left(2\lambda y\right)}{2\lambda }}.$$
(98)
Close to the throat $`\mathrm{exp}\left(2\lambda y\right)12\lambda y`$ and
$$\mu \left(\lambda \right)=\lambda ^2\frac{3MG}{\overline{r}^3\stackrel{~}{\kappa }^2}+\frac{9MG}{\overline{r}^3\stackrel{~}{\kappa }^2}\left[\frac{\overline{y}}{2\lambda }+\overline{y}^2\right].$$
(99)
The minimum of $`\mu \left(\lambda \right)`$ is reached for $`\overline{\lambda }=\left(\frac{9MG}{4\overline{r}^3\stackrel{~}{\kappa }^2}\overline{y}\right)^{\frac{1}{3}}`$ with the help of Eq.$`\left(\text{12}\right)`$, we can write
$$\mu \left(\overline{\lambda }\right)=3\left(\frac{9}{4}D\overline{y}\right)^{\frac{2}{3}}3D+3D\overline{y}^2,$$
(100)
where
$$D=\frac{MG}{\overline{r}^3\stackrel{~}{\kappa }^2}=2\frac{\overline{r}^2+b^2}{3\overline{r}^2+b^2}.$$
(101)
If we set the value of $`\overline{r}`$ equal to the surface gravity minimum location, then Eq.$`\left(\text{100}\right)`$becomes
$$\mu \left(\overline{\lambda }\right)=3\left(3\overline{y}\right)^{\frac{2}{3}}4+12\overline{y}^2,$$
(102)
which is zero for $`\overline{y}_c=\mathrm{.309\hspace{0.17em}15}`$ corresponding to $`\overline{\rho }_c=1.0956`$. This means that the unstable mode persists until the boundary radius $`\overline{\rho }`$ falls below $`\overline{\rho }_c`$, to be compared with the value $`\rho =1`$ of Refs..
## VIII Summary and Conclusions
Following Refs., in this paper we have computed the Casimir-like energy for a S-AdS space with a AdS space as a reference space. Due to the same asymptotic properties of these spaces and looking at the extended Kruskal S-AdS manifold at constant time, we have found that the classical contribution coming from boundaries disappears. According to Witten , since the energy is conserved and since boundary conditions are the same we can discuss the existence of an instability at zero temperature. A proof of instability at finite temperature has been given by Prestidge in Ref., based on conjectures of Hawking and Page . The zero temperature one-loop analysis shows a single negative mode in S wave which is interpreted as a clear signal of a decay form a false vacuum to a true one. This is also confirmed by one-loop stable modes contribution which shifts the energy minimum to the S-AdS space. Following the same procedure of Ref., we discover a critical radius $`r_c`$ below which the system becomes stable. In analogy with the flat space case, the appearance of this instability even at zero temperature is attributed to a neutral S-AdS black hole pair creation mediated by a three-dimensional S-AdS wormhole with the holes residing in different universes. The probability of creating such a pair and therefore the instability appearance at zero temperature is measured by
$$\mathrm{\Gamma }_{1\mathrm{hole}}=\frac{P_{\mathrm{S}\mathrm{AdS}}}{P_{\mathrm{AdS}}},$$
(103)
where
$$P\left|\mathrm{exp}\left(\mathrm{\Delta }E\right)\left(\mathrm{\Delta }t\right)\right|^2.$$
(104)
In spite of the evident analogy between the S-AdS and the Schwarzschild space it is not simple at this stage speculate on a possible foam-like structure composed by $`N`$ S-AdS coherent wormholes because the boundary reduction, which is fundamental to have the stabilization of the system in examination, is not of straightforward application due to the presence of the negative cosmological factor proportional to the square of the radius. However, if such a reduction mechanism could work $`we`$should discuss what is the meaning of
$$\mathrm{\Gamma }_{\mathrm{N}\mathrm{S}\mathrm{AdS}\mathrm{holes}}=\frac{P_{\mathrm{N}\mathrm{S}\mathrm{AdS}\mathrm{holes}}}{P_{\mathrm{AdS}}}\frac{P_{\mathrm{foam}}}{P_{\mathrm{AdS}}}$$
(105)
compared with
$$\mathrm{\Gamma }_{\mathrm{N}\mathrm{holes}}=\frac{P_{\mathrm{N}\mathrm{holes}}}{P_{\mathrm{flat}}}\frac{P_{\mathrm{foam}}}{P_{\mathrm{flat}}}.$$
(106)
## IX Acknowledgements
The author would like to thank the Referees for useful comments and suggestions, in particular for having brought to my attention the paper of Ref..
## A Kruskal-Szekeres coordinates for S-AdS spacetime
Before introducing the Kruskal-Szekeres type coordinates we recall that the S-AdS line element is defined as
$$ds^2=\left(1\frac{2MG}{r}+\frac{r^2}{b^2}\right)dt^2+\left(1\frac{2MG}{r}+\frac{r^2}{b^2}\right)^1dr^2+r^2d\mathrm{\Omega }^2.$$
(A1)
By looking at the $`(t,r)`$ coordinates one gets
$`ds^2=\left(1{\displaystyle \frac{2MG}{r}}+{\displaystyle \frac{r^2}{b^2}}\right)\left[dt^2dr^2\right]+r^2d\mathrm{\Omega }^2`$
$$=\left(1\frac{2MG}{r}+\frac{r^2}{b^2}\right)dvdu+r^2(u,v)d\mathrm{\Omega }^2.$$
(A2)
$`v=t+r^{}`$ is the ingoing radial null coordinate, $`u=tr^{}`$ is the outgoing radial null coordinate and
$$dr^{}=\frac{rb^2dr}{\left(r\overline{r}\right)\left(r^2+\overline{r}r+\overline{r}^2+b^2\right)}.$$
(A3)
The explicit integration gives $`2\kappa r^{}`$
$$=\mathrm{ln}\left|\frac{r}{\overline{r}}1\right|\frac{1}{2}\mathrm{ln}\left(\frac{r^2+\overline{r}r+\overline{r}^2+b^2}{\overline{r}^2+b^2}\right)+\frac{3\overline{r}^2+2b^2}{\overline{r}\sqrt{3\overline{r}^2+4b^2}}\mathrm{arctan}\left(\frac{2r\sqrt{3\overline{r}^2+4b^2}}{4\overline{r}^2+4b^2+\overline{r}r}\right),$$
(A4)
where $`\kappa `$ is the surface gravity defined in Eq.$`\left(\text{76}\right)`$. We now introduce Kruskal-Szekeres coordinates $`(U,V)`$ defined (for $`r>\overline{r}`$) by
$$U=e^{\kappa u}V=e^{\kappa v}$$
(A5)
with
$$UV=\mathrm{exp}\kappa \left(vu\right)=\mathrm{exp}\left(2\kappa r^{}\right)$$
(A6)
$$=\frac{\left(r/\overline{r}1\right)\sqrt{\overline{r}^2+b^2}}{\sqrt{r^2+\overline{r}r+\overline{r}^2+b^2}}\mathrm{exp}\left(\frac{3\overline{r}^2+2b^2}{\overline{r}\sqrt{3\overline{r}^2+4b^2}}\mathrm{arctan}\left(\frac{2r\sqrt{3\overline{r}^2+4b^2}}{4\overline{r}^2+4b^2+\overline{r}r}\right)\right)$$
(A7)
and
$$\frac{U}{V}=\mathrm{exp}\kappa \left(v+u\right)=\mathrm{exp}\left(2\kappa t\right)$$
(A8)
In terms of these coordinates Eq.$`\left(\text{A2}\right)`$ becomes
$$ds^2=\frac{\overline{r}\left(r^2+\overline{r}r+\overline{r}^2+b^2\right)^{\frac{3}{2}}}{\sqrt{\overline{r}^2+b^2}rb^2\kappa ^2}\mathrm{exp}\left(F\left(r\right)\right)dUdV+r^2(U,V)d\mathrm{\Omega }^2,$$
(A9)
where
$$F\left(r\right)=\frac{3\overline{r}^2+2b^2}{\overline{r}\sqrt{3\overline{r}^2+4b^2}}\mathrm{arctan}\left(\frac{2r\sqrt{3\overline{r}^2+4b^2}}{4\overline{r}^2+4b^2+\overline{r}r}\right).$$
(A10)
The only true singularities are at curves $`UV=1`$, where $`r=0`$. The region $`\left\{U<0,V>0\right\}`$ is the โoutside regionโ, the only region from which distant observers can obtain any information. The line $`V=0`$, where $`r=\overline{r}`$, is the โpast horizonโ; the line $`U=0`$ where also $`r=\overline{r}`$, is the โfuture horizonโ.
|
warning/0006/hep-th0006200.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Recently a model has been studied where new inflation is driven by a slow-rolling inflaton field, characterised by a quadratic potential, and incorporating radiative corrections within the context of supergravity. The so called $`\eta `$-problem is dealt with by radiative corrections to the inflaton mass $`m_\varphi ^2`$ which reduce its value from the Planck scale. A light inflaton field is confined at the origin by thermal effects naturally generating the initial conditions for a (last) stage of new inflation. Low powers of the inflaton dominate the potential during the era of observable inflation thus generating โquadraticโ inflation. The nice features of this model are that inflation can occur at the scale of supersymmetry breaking thus without having to invoke a new scale for inflation. Also the possibility of having electroweak scale inflation is realized without any extra difficulty. To implement this model a superpotential of the hybrid type containing two fields was used .
Here, we would like to explore the possibility of obtaining similar results in a more economical model with a single scalar field. For this an $`R`$-invariant superpotential is proposed in such a way that we can maintain the most important conclusions discussed previously in . In particular the $`R`$-symmetry of the superpotential restricts the powers that the inflaton can have. This forbids certain models which occur in but still maintaining others with low scales of inflation.
Analytical solutions can be worked out and a full description of the various quantities of interest during the inflationary era is given. In particular we find models which allow scales as low as the supersymmetry breaking scale of $`10^{10}`$ GeV or even the electroweak scale of $`10^3`$ GeV which could be relevant in the context of theories with submillimiter dimensions . We also find that the reheat temperature is not sufficiently high in general thus some other more efficient mechanism should be at work to attain higher reheat temperatures.
## 2 A Model for Low Scale Inflation
The model we propose to study is given by the following superpotential
$$W(\varphi )=\mathrm{\Delta }^2\varphi (1\frac{\kappa }{p+1}\frac{\varphi ^p}{\mathrm{\Delta }^q}),$$
(1)
and the Kรคhler potential
$$K(\varphi ,\varphi ^{})=\varphi \varphi ^{}+\frac{\mu }{4}(\varphi \varphi ^{})^2+\mathrm{},$$
(2)
where p, q are integer numbers and $`\mu `$ is a constant parameter with a value fixed by the inflationary constraints. The quantities $`\kappa `$, $`\mu `$, and $`\mathrm{\Delta }`$ have dimensions of $`M^{qp}`$, $`M^2`$, and $`M`$ respectively. From now on we will take $`MM_\mathrm{P}/\sqrt{8\pi }=1`$. The inflaton superfield $`\varphi `$ and $`\mathrm{\Delta }`$ have R-charges given by
$$R\varphi (\theta )=\frac{2}{n},R\mathrm{\Delta }^2=2\frac{2}{n}.$$
(3)
That is the superfield $`\varphi (\theta )`$ transforms
$$\varphi (\theta )\varphi ^{^{}}(\theta ^{^{}})=\mathrm{e}^{i\frac{2}{n}\alpha }\varphi (e^{i\alpha }\theta ),$$
(4)
where $`n`$ is a positive integer.
The form of Eqs.(1)-(4) has been studied before by Izawa and Yanagida where they consider a natural inflationary model in broken supergravity based on an R-symmetry. The new ingredient in our superpotential is the appearance of the scale of inflation $`\mathrm{\Delta }`$ in the higher dimension non-renormalizable terms. As has been discussed at lenght in , these higher order terms might arise as a result of integrating out heavy fields in the theory thus generating a mass scale $`M^{^{}}`$ in the denominator much less than the Planck scale. The scale $`M^{^{}}`$ can be associated with any of the scales in the theory in particular with the inflationary scale simply by writing $`M^p=\mathrm{\Delta }^qM^{pq}`$ (see Section 5). This avoids the introduction of yet another scale in the model and allows the interesting possibility of identifying the scale of inflation $`\mathrm{\Delta }`$ with that of supersymmetry breaking or even with the electroweak scale, of interest for theories with large extra dimensions . As has been shown before the factor $`1/\mathrm{\Delta }^q`$ in the higher order terms allows, in quadratic inflation, practically any scale of inflation.
The scale $`\mathrm{\Delta }`$ can be though as due to the presence of a composite superfield which condenses when the (gaugino) interaction becomes strong at the scale $`\mathrm{\Delta }`$, breaking the $`U(1)_R`$ or $`Z_n`$ symmetry of the model. This R-symmetry specifies the superpotential and imposes the following relation between p and q
$$p=\frac{q}{2}(n1).$$
(5)
It is the presence of the scale $`\mathrm{\Delta }`$ through the factor $`\mathrm{\Delta }^q`$ in Eq.(1) which allows to have low scale inflation as shown below. Also, the $`\mu `$-parameter appearing in the Kรคhler potential Eq.(2) enters in the mass term for the inflaton $`m_\varphi ^2\mu \mathrm{\Delta }^4`$. No other contributions to $`m_\varphi ^2`$ occur in the tree level potential . To show this let us consider the supergravity potential
$$V=\mathrm{exp}\left(K\right)\left[F^A(K_A^B)^1F_B3|W|^2\right]+\mathrm{D}\mathrm{terms},$$
(6)
where
$$F_A\frac{W}{\mathrm{\Phi }^A}+\left(\frac{K}{\mathrm{\Phi }^A}\right)W,\left(K_A^B\right)^1\left(\frac{^2K}{\mathrm{\Phi }^A\mathrm{\Phi }_B^{}}\right)^1.$$
(7)
For small field values we can expand $`V`$ so that
$$V\mathrm{\Delta }^4(1\mu \varphi ^2+\mu ^{^{}}\varphi ^42\kappa \frac{\varphi ^p}{\mathrm{\Delta }^q}+\kappa ^2\frac{\varphi ^{2p}}{\mathrm{\Delta }^{2q}}+\mathrm{}),$$
(8)
where $`\mu ^{^{}}=2\frac{7}{4}\mu +\mu ^2.`$ Since $`\mathrm{\Delta }1`$ the $`\varphi ^4`$ term is much less than $`\varphi ^p/\mathrm{\Delta }^q`$ for $`p=4`$. For $`p>4`$, $`\varphi ^4\varphi ^p/\mathrm{\Delta }^q`$ whenever $`\varphi \mathrm{\Delta }^{\frac{q}{p4}}`$ which is always the case in the examples of interest we study below. When $`\varphi `$ is much less than one, higher order terms in $`\varphi `$ are negligible, and have been omitted in Eq.(8). In this case we can work with the simpler expression
$$\frac{V}{\mathrm{\Delta }^4}=\left(1\kappa \frac{\varphi ^p}{\mathrm{\Delta }^q}\right)^2\mu \varphi ^2,$$
(9)
which is practically indistinguishable from the full supergravity potential Eq.(6) all the way to the global minimum.
## 3 Analytical Solutions
Here we obtain closed form expressions for the relevant quantities involved in the inflationary era. We are assuming that the radiative corrections to the inflaton mass $`\mathrm{ln}\varphi `$ are already included in the parameter $`\mu `$ and we take $`\mu `$ fixed by its value at $`\varphi _\mathrm{H}`$ (where the subscript $`\mathrm{H}`$ denotes the epoch at which a fluctuation of wavenumber $`k`$ crosses the Hubble radius $`H^1`$ during inflation). This is not a great sin since the $`\mathrm{ln}\varphi `$ corrections change very slowly from $`\varphi _\mathrm{H}`$ to the end of inflation at $`\varphi _e`$ and it turns out to be a very good approximation to consider $`\mu `$ as a constant. The advantage of doing this is that we can obtain closed form solutions. The parameter $`\mu `$ can take positive or negative values. In particular when $`\mu <0`$ there is a maximum at
$$\varphi _{max}\left(\frac{\mu \mathrm{\Delta }^q}{\kappa p}\right)^{\frac{1}{p2}},$$
(10)
when $`\mu 0,\varphi _{max}0`$ as it should. In this case the $`\mu \varphi ^2`$-term dominates $`V(\varphi )`$ in the interval $`0\varphi \varphi _{max}`$. Inflation for $`\varphi >\varphi _{max}`$ requires the participation of both $`\mu \varphi ^2`$ and $`2\kappa \varphi ^p/\mathrm{\Delta }^q`$ with the last term dominating during inflation. Thus we cannot talk about โquadraticโ inflation when $`\mu <0`$, this can only occur for positive $`\mu `$. The following expressions, however, are valid for any $`\mu `$.
1) The end of inflation. In the models under consideration inflation is generated while $`\varphi `$ rolls to larger values. The end of inflation occurs at $`\varphi =\varphi _e`$ when the slow roll conditions are violated. The slow-roll conditions are upper limits on the normalised slope and curvature of the potential:
$$ฯต\frac{1}{2}\left(\frac{V^{}}{V}\right)^2\gamma ,|\eta |\left|\frac{V^{\prime \prime }}{V}\right|\gamma .$$
(11)
The potential determines the Hubble parameter during inflation as $`H_{\mathrm{inf}}\dot{a}/a\sqrt{V/3M^2}`$. Inflation ends (i.e. $`\ddot{a}`$, the acceleration of the cosmological scale factor, changes sign from positive to negative) when $`ฯต`$ and/or $`|\eta |`$ become of $`๐ช(\gamma )`$. This occurs at $`V^{\prime \prime }(\varphi )\gamma `$, where $`\gamma =๐ช(1).`$ Thus we have
$$\varphi _e\left(\frac{(\gamma 2\mu )\mathrm{\Delta }^q}{2\kappa p(p1)}\right)^{\frac{1}{p2}}.$$
(12)
2) Scalar density perturbations. The adiabatic scalar density perturbation generated through quantum fluctuations of the inflaton is
$$\delta _\mathrm{H}^2(k)=\frac{1}{150\pi ^2}\frac{V_\mathrm{H}}{ฯต_\mathrm{H}},$$
(13)
where the subscript $`\mathrm{H}`$ denotes the epoch at which a fluctuation of wavenumber $`k`$ crosses the Hubble radius $`H^1`$ during inflation, i.e. when $`aH=k`$. (We normalise $`a=1`$ at the present epoch, when the Hubble expansion rate is $`H_0100h`$ km s<sup>-1</sup>Mpc<sup>-1</sup>, with $`h0.50.8`$). The COBE observations of anisotropy in the cosmic microwave background on large angular-scales require
$$\delta _{\mathrm{COBE}}1.9\times 10^5,$$
(14)
on the scale of the observable universe ($`k_{\mathrm{COBE}}^1H_0^13000h^1`$ Mpc). In addition, the COBE data fix the spectral index, $`n_\mathrm{H}(k)1+\mathrm{d}\delta _\mathrm{H}^2(k)/\mathrm{d}\mathrm{ln}k=16ฯต_\mathrm{H}+2\eta _\mathrm{H}`$, on this scale:
$$n_{\mathrm{COBE}}=1.2\pm 0.3.$$
(15)
Solving Eq. (13) we find
$$\varphi _\mathrm{H}^{p1}+\frac{\mu \mathrm{\Delta }^q}{\kappa p}\varphi _\mathrm{H}\frac{\mathrm{\Delta }^{q+2}}{2\kappa pA_\mathrm{H}}=0,$$
(16)
where $`A_\mathrm{H}\sqrt{75}\pi \delta _\mathrm{H}.`$ This equation determines $`\mathrm{\Delta }`$ once $`\varphi _\mathrm{H}`$ is determined.
3) Number of e-folds. The number of e-folds from $`\varphi _\mathrm{H}`$ to the end of inflation at $`\varphi _e`$ is
$`N_\mathrm{H}`$ $``$ $`{\displaystyle _{\varphi _\mathrm{H}}^{\varphi _\mathrm{e}}}{\displaystyle \frac{V(\varphi )}{V^{}(\varphi )}}d\varphi {\displaystyle \underset{\varphi _H}{\overset{\varphi _e}{}}}๐\varphi {\displaystyle \frac{1}{2\mu \varphi +2\kappa p\varphi ^{p1}/\mathrm{\Delta }^q}}`$ (17)
$`=`$ $`{\displaystyle \frac{1}{2\mu (p2)}}\mathrm{ln}\left({\displaystyle \frac{1+\frac{\mu \mathrm{\Delta }^q}{\kappa p\varphi _e^{p2}}}{1+\frac{\mu \mathrm{\Delta }^q}{\kappa p\varphi _\mathrm{H}^{p2}}}}\right).`$
Solving for $`\varphi _\mathrm{H}`$ gives
$$\varphi _\mathrm{H}=\left(\frac{\mu \mathrm{\Delta }^q}{\kappa p(1(1+\frac{2\mu (p1)}{\gamma 2\mu })e^{2\mu (p2)N_\mathrm{H}})}\right)^{\frac{1}{p2}}B\mathrm{\Delta }^{\frac{q}{p2}}.$$
(18)
Finally substituting in Eq.(16) and simplifying we obtain the required solution for $`\mathrm{\Delta }`$
$$\mathrm{\Delta }=\left(2\kappa pA_\mathrm{H}(B^{p1}+\frac{\mu B}{\kappa p})\right)^{\frac{p2}{2(p2)q}}.$$
(19)
4) Spectral Index. We now readily obtain a form for the spectral index
$`n_\mathrm{H}`$ $``$ $`1+2V^{\prime \prime }(\varphi _\mathrm{H})`$ (20)
$``$ $`14\mu 4\kappa p(p1)B^{p2}.`$
5) Reheat temperature. One obvious effect of lowering the scale of the inflationary potential is the lowering of the reheat temperature. At the end of inflation the oscillations of the inflaton field would make it decay thus reheating the universe. The couplings of the inflaton to some other bosonic $`\chi `$ or fermionic $`\psi `$ MSSM fields occur due to terms $`\frac{1}{2}g^2\varphi ^2\chi ^2`$ or $`h\overline{\psi }\psi \varphi `$, respectively. These couplings induce decay rates of the form
$$\mathrm{\Gamma }(\varphi \chi \chi )=\frac{g^4\varphi _0^2}{8\pi m_\varphi },\mathrm{\Gamma }(\varphi \overline{\psi }\psi )=\frac{h^2m_\varphi }{8\pi }$$
(21)
where $`\varphi _0`$ is the value of $`\varphi `$ at the minimum of the potential
$$\varphi _0\left(\frac{\mathrm{\Delta }^q}{\kappa }\right)^{\frac{1}{p}},$$
(22)
and $`m_\varphi `$ is the inflaton mass given by
$$m_\varphi \sqrt{2}p\kappa ^{\frac{1}{p}}\mathrm{\Delta }^{2\frac{q}{p}}.$$
(23)
A maximum value for the decay is obtained when $`m_{\chi ,\psi }m_\varphi `$. In this case we find
$$\mathrm{\Gamma }\frac{m_\varphi ^3}{8\pi \varphi _0^2}$$
(24)
The reheat temperature at the beginning of the radiation-dominated era is thus
$$T_{rh}\left(\frac{90}{\pi ^2g_{}}\right)^{\frac{1}{4}}min(\sqrt{H(\varphi _e)},\sqrt{\mathrm{\Gamma }})\left(\frac{30}{\pi ^2g_{}}\right)^{\frac{1}{4}}min(\mathrm{\Delta },\left(\frac{3}{8\pi ^2}\right)^{\frac{1}{4}}p^{\frac{3}{2}}\kappa ^{\frac{5}{2p}}\mathrm{\Delta }^{3\frac{5q}{2p}})$$
(25)
where $`g_{}`$ is the number of relativistic degrees of freedom which for the MSSM equal $`915/4`$.
6) Quantum fluctuations. The value $`\varphi _\mathrm{H}`$ at the beginning of the last $`N_\mathrm{H}`$ e-folds of inflation should exceed the quantum fluctuations of the inflaton $`\delta \varphi \frac{H}{2\pi }\frac{\mathrm{\Delta }^2}{2\pi \sqrt{3}}`$. From Eqs.(18) and (19) we can impose the following condition
$$\frac{\delta \varphi }{\varphi _\mathrm{H}}\frac{\mathrm{\Delta }^{\frac{2(p2)q}{p2}}}{2\pi \sqrt{3}B}(\mu +\kappa pB^{p2})\times 10^41.$$
(26)
Typically $`B10^2`$ and the condition Eq.(26) is easily verified by the models we are interested in.
## 4 Numerical Results
The number $`N_\mathrm{H}`$ of $`e`$-folds of the present horizon is given by
$$N_\mathrm{H}67+\frac{1}{3}\mathrm{ln}H+\frac{1}{3}\mathrm{ln}T_{rh}67+\frac{1}{3}\mathrm{ln}\left(\left(\frac{10}{3\pi ^2g_{}}\right)^{\frac{1}{4}}min(\mathrm{\Delta }^3,\left(\frac{3}{8\pi ^2}\right)^{\frac{1}{4}}p^{\frac{3}{2}}\kappa ^{\frac{5}{2p}}\mathrm{\Delta }^{5\frac{5q}{2p}})\right).$$
(27)
Solving this equation consistently with Eq.(19) we can obtain a representative set of values for the various quantities of interest during inflation. A sample is given in Table 1.
We now plot in $`Fig.1`$ the inflaton potential $`V(\varphi ,\alpha )`$ as a function of $`\varphi `$ and the phase $`\alpha `$. In $`Figs.2a,2b`$, and $`2c`$ we plot the scale of inflation $`\mathrm{\Delta }`$, the reheating temperature $`T_{rh}`$, and the spectral index $`n_\mathrm{H}`$ as functions of the parameter $`\mu `$, respectively. Finally $`Fig.2d`$ shows the behaviour of the spectral index as a function of the number $`N_\mathrm{H}`$ of e-folds of inflation from the end of inflation. All of these figures are for the case $`(p,q)=(4,2)`$. Similar behaviour is found in the other $`(p,q)`$ cases.
## 5 Comparison with related work
In the models studied in and further elaborated in quadratic inflation is implemented through a hybrid mechanism with the participation of two fields. A linear term in a field $`Y`$ follows if $`Y`$ carries non-zero $`R`$-symmetry charge under an unbroken $`R`$-symmetry. The inflaton $`\varphi `$ is a singlet under the $`R`$-symmetry but carries a charge under a discrete $`Z_p`$ symmetry. Then the superpotential has the form
$$W=\left(\mathrm{\Delta }^2\frac{\varphi ^p}{M^{p2}}\frac{\varphi ^{2p}}{M^{2p2}}\mathrm{}\right)Y.$$
(28)
This gives rise to the potential
$$V=\left(\mathrm{\Delta }^2\frac{\varphi ^p}{M^{p2}}\frac{\varphi ^{2p}}{M^{2p2}}\mathrm{}\right)^2,$$
(29)
displaying the possibilities of ending inflation. There are also terms involving $`Y`$ which are dropped as they do not contribute to the vacuum energy since $`Y`$ does not acquire a vacuum expectation value. The scale $`M^{}`$ denotes new physics below the Planck scale and we can write $`M^{p2}=\mathrm{\Delta }^qM^{pq2}`$ to take into account the possibility that the scale associated with the higher dimension operators may be below the Planck scale.
In the present model there is only a single scalar field with a superpotential determined by the $`R`$-symmetry as shown in Section 2. As a consequence of this symmetry some $`(p,q)`$ models which occur in , (for example $`(p,q)=(4,3)`$) are not allowed here. It is therefore interesting that most of the results and conclusions of are still maintained.
Other studies of quadratic inflation have concentrated on the case where radiative corrections make the potential develop a maximum near the origin, from which the inflaton rolls either away from the origin or towards it , and inflation ends through a hybrid mechanism.
There is also related work with a superpotential (in our notation)
$$W(\varphi )=\mathrm{\Delta }^2\varphi \frac{\kappa }{n}\varphi ^n,$$
(30)
and the Kรคhler potential
$$K(\varphi ,\varphi ^{})=\varphi \varphi ^{}+\frac{\mu }{4}(\varphi \varphi ^{})^2+\mathrm{},$$
(31)
where $`\varphi `$ and $`\mathrm{\Delta }`$ have R-charges as in Eq.(3). However the fact that the factor $`\mathrm{\Delta }^q`$ appearing in our Eq.(1) is not present in Eq.(30) above eliminates the possibility of having low scales for inflation (in the lowest scale allowed is $`๐ช(10^{12}GeV)).`$
## 6 Conclusions
We have studied a model of inflation where low inflationary scales are allowed without having to introduce unnatural values for the parameters involved. The model is defined in terms of a single scalar field, the inflaton, and the term driving (new) inflation is quadratic in $`\varphi `$. The end of inflation due to higher order non-renormalisable terms. Radiative corrections to the inflaton mass reduce $`m_\varphi ^2`$ from its natural value at the Planck scale. For a light inflaton thermal initial conditions can naturally place the inflaton at the origin, initiating a (last) stage of new inflation. A quadratic parameterisation of the inflationary potential allows low values for the inflationary scale $`\mathrm{\Delta }`$. One can have $`\mathrm{\Delta }10^{10}\mathrm{GeV}`$, the supersymmetry breaking scale in the hidden sector or the electroweak scale $`\mathrm{\Delta }10^3\mathrm{GeV}`$ which could be relevant in the context of theories with submillimeter dimensions. The well justified assumption that the inflaton mass parameter $`m_\varphi ^2\mu \mathrm{\Delta }^4`$ remains practically constant during inflation allows analytical closed form expressions for all the relevant quantities.
## 7 Acknowledgements
G.G. would like to thank G.G. Ross and S. Sarkar for useful discussions. This work was supported by the projects PAPIIT IN110200, and Conacyt 32415-E.
|
warning/0006/cond-mat0006188.html
|
ar5iv
|
text
|
# Non trivial overlap distributions at zero temperature
##
## I Introduction
In recent times there has been a wide interest in the behavior spin glasses with Gaussian couplings at zero temperature.
Some of the reasons for this interest are the following:
* The energies are continuous variables and the ground state is unique. It is also natural to suppose (although it is far from being proved) that the limits $`T0`$ and $`N\mathrm{}`$ do commute and therefore the shape of the energy landscape is similar to that of the free energy landscape at non-zero temperature (for a discussion of this point see ).
* Working at zero temperature avoids completely the possibility that the temperature used is too near to the critical point.
* Technical progresses has been done in the algorithm for finding the ground state and it is now possible to studies three dimensional systems up to $`14^3`$ spins .
In this framework it has been suggested that a possible test of the applicability of the Replica Symmetry Breaking (RSB) scenario is the study of the overlap of the ground state of two systems whose total Hamiltonian differs by a quantity of order 1 .
Let us consider a simple case. We have a first system with Hamiltonian $`H_0(\sigma )`$ and its ground state is given by $`\tau _i`$. We now consider a second system whose Hamiltonian is
$$H_1(\sigma )=H_0(\sigma )+ฯตH_\tau (\sigma ).$$
(1)
Three quite simple choices of $`H_\tau (\sigma )`$ are:
$$H_\tau (\sigma )=q(\sigma ,\tau ),H_\tau (\sigma )=q^2(\sigma ,\tau ),H_\tau (\sigma )=q_l(\sigma ,\tau ),$$
(2)
where the overlap $`q`$ and the link overlap $`q_l`$ are given by
$`q(\sigma ,\tau )=N^1{\displaystyle \underset{i}{}}\sigma _i\tau _i,`$ (3)
$`q_l(\sigma ,\tau )=N_l^1{\displaystyle \underset{i,k}{}}\sigma _i\tau _i\sigma _k\tau _k,`$ (4)
where the sum is done over all the nearest neighbor pairs $`(i,k)`$ in a short range model or over all the pairs in the SK model ($`N_l`$ being the total number of pairs in this sum). The third possibility has been actually used by .
In presence of quenched disorder, the the value of the overlap among the ground states of $`H_0`$ and $`H_1`$ can be sample dependent. This observation can be used tp the starting point for investigating possible RSB in the three dimensional Edwards-Anderson model. The question we address in this paper, is the computation of the probability distribution induced by the random couplings of $`q`$ or of $`q_l`$ among the two ground states, in the hierarchical RSB scenario.
Obviously the choice $`H_q(\sigma )=q`$ is interesting only in presence of a magnetic field which breaks the symmetry $`\sigma (i)\sigma (i)`$, otherwise we would get that $`\sigma (i)=\tau (i)`$ for positive non zero $`ฯต`$ and $`q=1`$. The second choice is more interesting at zero magnetic field, but it is slightly harder to implement numerically, because its non local nature. The third choice is however equivalent to the first one in the SK model, where is known that $`q_l=q^2`$ apart from corrections that vanishes when the number of spins goes to infinity. In short range models, it is possible (as suggested by the principle of replica equivalence ) that with probability one when the volume goes to infinity $`q_l=f(q^2)`$, where $`f`$ is a function that can be determined numerically and which should be not too far from
$$f(q^2)=A+(1A)q^2.$$
(5)
It is evident that for finite $`ฯต`$ the perturbation is of order 1 and it quite interesting that if replica symmetry is broken the function $`P(q)`$ is non trivial at $`ฯต0`$. Let us define as $`E_{gs}(0)+\mathrm{\Delta }(q)`$ the energy of the first excited state of the Hamiltonian $`H_0`$ with an overlap $`q`$ with the ground state. The ground state of the Hamiltonian $`H_1`$ is
$$E_{gs}(ฯต)=E_{gs}(0)+\underset{0q1}{\mathrm{min}}\{\mathrm{\Delta }(q)+ฯตg(q)\}$$
(6)
where $`g(q)=q,q^2,q_l`$. The main achievement of this paper will be the computation in section II and III of the joint probability distribution of $`\mathrm{\Delta }`$ and $`q`$ for which the minimum is attained, for arbitrary RSB trees. In section IV we will give some example in mean field models, while in section V we show the result of a numerical computation for directed polymers in random media in 1+1 dimension.
## II Replica symmetry breaking
The computation presented in this note could be done in two different ways:
* Using the replica formalism to compute the partition function of the perturbed Hamiltonian $`H_1`$.
* Exploiting directly the information coming from replica symmetry breaking on the probability distribution of the lowest lying states and doing a pure probabilistic computation.
The first alternative leads to some apparently messy combinatorial analysis so that we have decided to follow the second alternative. In this case the computation is physically instructive.
In this section we will recall, without proof, some known results about replica symmetry breaking and also find some new consequences of those results. Let us assume that replica symmetry is broken in the system we consider and its breaking is characterized by a function $`x(q,T)`$ such that in the low temperature limit
$$x(q,T)=Ty(q)+O(T^2),$$
(7)
where the function $`y(q)`$ may be singular at $`q=1`$ (in the SK model it diverges as $`(1q)^{1/2}`$ near $`q=1`$) .
The space of lower lying configurations is organized in a rather complex way.
### A One step replica symmetry breaking
In this case only two values of the overlap are allowed ($`q_0`$ and 1), i.e. all different minima have a mutual overlap equal to $`q_0`$. If we call $`R`$ a reference total energy, which depends on the choice of the systems, i.e. on the variables $`J`$, the probability to find a configuration in the interval $`(E,E+dE)`$ is given by
$$\nu _0(E|R)\mathrm{exp}(y_0(ER)).$$
(8)
We notice that configurations which differs by a number of spin flips which remains finite when the volume goes to infinity are identified.
This well known results has the consequence that the probability distribution of the ground state energy $`E_0`$ is given by the Gumbel law
$$\mu _0(E_0|R)=\mathrm{exp}(y_0(E_0R))\mathrm{exp}(A_0\mathrm{exp}(y_0(E_0R)),$$
(9)
with $`A_0=1/y_0`$. This formula is easily understood noticing that the probability that there are no configurations for $`E^{}<E`$ is given by
$$\mathrm{exp}(_{\mathrm{}}^EdE^{}\mathrm{exp}(y_0(E^{}R)))=\mathrm{exp}(y_0^1\mathrm{exp}(y_0(ER)).$$
(10)
In the same way we obtaining that the probability of having a ground state at $`E_0`$ and the first excited configuration at $`E_1`$ is given by:
$`P(E_0,E_1|R)`$ $`=`$ $`\nu _0(E_0|R)\mu _0(E_1|R),`$ (11)
$`=`$ $`\mathrm{exp}(y_0(E_0R))\mathrm{exp}(y_0(E_1R))\mathrm{exp}(y_0^1\mathrm{exp}(y_0(E_1R)).`$ (12)
Finally if we define $`\mathrm{\Delta }_0=E_1E_0`$, the probability distribution of $`\mathrm{\Delta }_0`$, integrated over $`E_0`$ and $`E_1`$, one finds the simple result
$$P(\mathrm{\Delta }_0)=y_0^1\mathrm{exp}(y_0\mathrm{\Delta }_0),$$
(13)
for positive $`\mathrm{\Delta }_0`$, the probability being obviously zero for negative $`\mathrm{\Delta }_0`$.
### B Many level replica symmetry breaking
Let us in this section generalize the computation for an arbitrary RSB tree (see figure). As customary, we will first consider a tree with $`k`$ levels and at the end we will generalize the result to the continuous branching limit. The construction of the tree has been described many times, and we only repeat it briefly to fix the notation. At each node of the tree at the level $`l`$ it is assigned an energy $`E_l`$, which is chosen in such a way that the number of nodes with energy in the interval $`(E_l,E_l+dE_l)`$ branching from a node with energy $`E_{l1}`$ is a Poisson variable with average equal to $`\mathrm{exp}(y_l(E_lE_{l1}))\mathrm{d}E_l`$. We will consider the case in which for all $`l`$, $`y_{l+1}>y_l`$, which will be a necessary condition of convergence of the integrals that appear in the computation. The root energy $`E_0=R`$ is the reference energy of the previous section. We call $`l`$-clusters, the set of branches which coincide at the $`l`$-th level of the tree.
Let us define for any $`l=k1,\mathrm{},1`$ the first $`l`$-excited state, as the first excited state which is in the same $`l`$-cluster as the ground state, but in a different $`l+1`$-cluster, and denote its energy $`E_{gs}+\mathrm{\Delta }_l`$. In this section we compute the joint probability distribution of all the $`l`$-gaps $`\mathrm{\Delta }_l`$. We will get this quantity by first computing $`P_k(E_{gs},E_{gs}+\mathrm{\Delta }_k,\mathrm{},E_{gs}+\mathrm{\Delta }_1|E_0)`$ and then integrating over $`E_{gs}`$. In that computation we make use of the following properties:
1. The probability of a ground state $`\mu _l(E|E_l)`$ of an $`l`$-cluster (we call it an $`l`$ ground state) is given by
$$\mu _l(E|E_l)=\mathrm{exp}\left(y_{l+1}(EE_l)+A_l\mathrm{e}^{y_{l+1}(EE_l)}\right).$$
(14)
where the $`A_l`$ are a positive constants whose value could be easily compute, but we will not need. For $`l=k`$ the formula was derived in the previous section. Let us now proceed by induction supposing that the formula holds for $`l+1`$ and show that it holds for $`l`$. Under the induction hypothesis, we find that the number of $`l+1`$-ground states with energy $`E`$ in an $`l`$ cluster is given by
$`\nu _{l+1}(E|E_l)\mathrm{d}E_l`$ $`=`$ $`{\displaystyle ๐E_{l+1}\mathrm{e}^{y_{l+1}(E_{l+1}E_l)}\mu _{l+1}(E|E_{l+1})}`$ (15)
$`=`$ $`const\times \mathrm{e}^{y_{l+1}(EE_l)}\mathrm{d}E_l`$ (16)
from which we immediately find that the distribution of the $`l`$ ground state is given by (14) exploiting the reasonings of the previous sub-section.
2. The joint probability $`P_k(E_{gs},E_{gs}+\mathrm{\Delta }_k,\mathrm{},E_{gs}+\mathrm{\Delta }_1|E_k,E_{k1},\mathrm{},E_0)`$ can be written as:
$`P_k(E_{gs},E_{gs}+\mathrm{\Delta }_{k1},\mathrm{},E_{gs}+\mathrm{\Delta }_0|E_{k1},\mathrm{},E_0)=`$ (17)
$`\nu _k(E_{gs}|E_{k1})\mu _k(E_{gs}+\mathrm{\Delta }_{k1}|E_{k1})\mu _{k1}(E_{gs}+\mathrm{\Delta }_{k2}|E_{k2})\times \mathrm{}\times \mu _1(E_{gs}+\mathrm{\Delta }_0|E_0)`$ (18)
from which we get:
$`P_k(E_{gs},E_{gs}+\mathrm{\Delta }_k,\mathrm{},E_{gs}+\mathrm{\Delta }_0|E_0)=`$ (19)
$`{\displaystyle }\mathrm{d}E_{k1}\mathrm{}\mathrm{d}E_1\nu _k(E_{gs}|E_{k1})\mu _k(E_{gs}+\mathrm{\Delta }_k|E_{k1})\mathrm{e}^{y_{k1}(E_{k1}E_{k2})}\mu _{k1}(E_{gs}+\mathrm{\Delta }_{k2}|E_{k2})\mathrm{e}^{y_{k2}(E_{k2}E_{k3})}\times `$ (20)
$`\mathrm{}\times \mu _2(E_{gs}+\mathrm{\Delta }_2|E_1)\mathrm{e}^{y_1(E_1E_0)}\mu _1(E_{gs}+\mathrm{\Delta }_1|E_0)`$ (21)
3. A detailed computation shows that
$$dE_{k1}\nu _k(E_{gs}|E_{k1})\mu _k(E_{gs}+\mathrm{\Delta }_k|E_{k1})\mathrm{e}^{y_{k1}(E_{k1}E_{k2})}=(y_ky_{k1})\mathrm{e}^{(y_ky_{k1})\mathrm{\Delta }_k}\nu _{k1}(E_{gs}|E_{k2}).$$
(22)
This allows to integrate all the $`E_l`$ ($`l=k1,\mathrm{},1`$) telescopically, and obtain
$$P_k(E_{gs},E_{gs}+\mathrm{\Delta }_k,\mathrm{},E_{gs}+\mathrm{\Delta }_1|E_0)=\underset{i=2}{\overset{k}{}}\left((y_iy_{i1})\mathrm{e}^{(y_iy_{i1})\mathrm{\Delta }_i}\right)\nu _1(E_{gs}|E_0)\mu _1(E_{gs}+\mathrm{\Delta }_1|E_0)$$
(23)
4. We can finally integrate $`E_{gs}`$ as in the previous section and get that the gapsโ probability distribution is:
$$P(\mathrm{\Delta }_k,\mathrm{},\mathrm{\Delta }_1)=\underset{i=1}{\overset{k}{}}\left((y_iy_{i1})\mathrm{e}^{(y_iy_{i1})\mathrm{\Delta }_i}\right)$$
(24)
having defined $`y_0=0`$.
We can now consider the continuum branching limit, in which the branches can be indexed by the value of q, or by any monotonically increasing function of $`q`$. $`y_iy(q)`$, $`y_iy_{i1}y^{}(q)dq`$ where the previous formula reduces to
$$P(\{\mathrm{\Delta }(q)\})=\mathrm{exp}\left(_0^1๐qy^{}(q)\mathrm{\Delta }(q)\right)\underset{q=0}{\overset{1}{}}y^{}(q)dq,$$
(25)
where, if $`y_1y(0)0`$, we make the convention that $`y^{}(0)=y(0)\delta (q)`$.
## III The overlap probability distribution
We are finally in the position to compute the joint distribution of the gap and of the overlap. For a given sample, as we said, the difference among the ground state energy of the Hamiltonians $`H_1`$ and $`H_0`$ is given by the
$$E_{gs}(ฯต)E_{gs}(0)=\underset{0q1}{\mathrm{min}}\mathrm{\Delta }(q)+ฯตg(q).$$
(26)
Noticing that the indexing of the tree could be done by the function $`g(q)`$ itself, we can concentrate here to case $`g(q)=|q|`$, and consider only positive overlaps. All the other cases can be obtained by this one via a simple change of variable. Let us call $`\mathrm{\Delta }`$ and $`q`$ the arguments of the minimum in (26). Notice that if we want $`q1`$, $`\mathrm{\Delta }`$ has to verify the inequality $`\mathrm{\Delta }ฯต(1q)`$, which expresses the fact that $`\mathrm{\Delta }(1)`$ is always equal to zero.
We get the probability $`P(\mathrm{\Delta },q)`$ integrating in formula (25) all the $`\mathrm{\Delta }(q^{})`$ ($`q^{}q`$) with the condition $`\mathrm{\Delta }(q^{})\mathrm{min}\{0,\mathrm{\Delta }+ฯต(qq^{})\}`$, so that, if $`q^{}>\mathrm{\Delta }/ฯต+q`$ the integration over $`\mathrm{\Delta }(q^{})`$ will contribute with a one, while, in the opposite case $`q^{}<\mathrm{\Delta }/ฯต+q`$ it contributes with the factor $`\mathrm{exp}(ฯตdq^{}y^{}(q^{})(\mathrm{\Delta }/ฯต+qq^{}))`$. We finally get:
$$P(\mathrm{\Delta },q)=\theta (1q\mathrm{\Delta }/ฯต)y^{}(q)\mathrm{exp}\left(ฯต_0^{q+\mathrm{\Delta }/ฯต}๐q^{}(y(q^{})y(0))\right)+\delta (\mathrm{\Delta })\delta (q1)\mathrm{exp}(ฯต\chi ),$$
(27)
where we defined $`\chi =y(0)+_0^1๐qy(q)`$. The factor $`y^{}(q)`$ that multiply the exponential comes from the only $`\mathrm{\Delta }`$ which has remained unintegrated.
Let us notice that the formula depends on $`\mathrm{\Delta }`$ only in the combination $`\mathrm{\Delta }/ฯต+q`$.
Integrating over $`\mathrm{\Delta }`$ we get, for the overlap probability
$$P(q)=\delta (q1)\mathrm{exp}(ฯต\chi )+ฯตy^{}(q)_q^1๐q^{}\mathrm{exp}\left(ฯต_0^q^{}๐q^{\prime \prime }(y(q^{\prime \prime })y(0))\right)$$
(28)
It is also interesting to study the probability distribution of $`w=\mathrm{\Delta }/ฯต+q`$. Integrating over $`\mathrm{\Delta }`$ and $`q`$ for fixed $`w`$ with the condition $`0qw`$ we find the remarkable formula:
$$P(w)=\theta (1w)ฯต(y(w)y(0))\mathrm{exp}\left(ฯต_0^wdq(y(q)y(0))\right)+\delta (w1)\mathrm{exp}(ฯต\chi ).$$
(29)
The primitive of this function has a very simple dependence on $`y`$ and $`ฯต`$. If we define $`Q(w)=_w^1dw^{}P(w^{})`$ we find:
$$Q(w)=\mathrm{exp}\left(ฯต_0^wdq(y(q)y(0))\right)$$
(30)
which is particularly well suited for the extraction of the function $`y(q)`$ from numerical simulations. Notice that eq. (30) implies that $`ฯต^1\mathrm{ln}(Q(w))`$ is $`ฯต`$-independent, which is a non-trivial result.
## IV Examples
In this section we show how our formula looks like in some cases. The first example we make is the one of spherical models. These models are defined by a random Gaussian Hamiltonian $`H(\sigma )`$ ($`\sigma =\{\sigma _1,\mathrm{},\sigma _N\}`$), which have correlation function
$$\overline{H(\sigma )H(\tau )}=Nf(q(\sigma ,\tau ))$$
(31)
and the spins are subject to the spherical constraint $`_i\sigma _i^2=N`$. In these models, the function $`y(q)`$ is temperature independent and equal to
$$y(q)=1/2f^{\prime \prime \prime }(q)/(f^{\prime \prime }(q)^{3/2}).$$
(32)
This last equation makes sense only if the resulting function $`y(q)`$ is an increasing function of $`q`$. This in particular happens for $`f(q)=1/2(q^2+aq^p)`$ if $`p4`$ and $`a`$ small enough, where we find
$$y(q)=\frac{a}{\sqrt{2}}\frac{1}{(p3)!}\frac{q^{p3}}{\left(2+\frac{a}{(p2)!}q^{p2}\right)^{3/2}},$$
(33)
while
$$Y(w)_0^wdqy(q)=1\frac{1}{\sqrt{1+\frac{aq^{p2}}{2(p2)!}}}.$$
(34)
In figure 2 we show the function $`Y(w)`$ and the function $`Q(w)`$ for various values of $`ฯต`$ in the case $`p=5`$, $`a=0.3`$. In figure 3 we show the function $`P(q)`$ for the same values of the parameters.
In the case of the SK model the function $`y(q)`$ at low temperature, has be estimated in using the so-called PaT approximation, and displays a square root divergence at $`q=1`$, while starting linearly at $`q=0`$ (a best fit of the form $`y(q)=\frac{aq+bq^2}{\sqrt{1q}}`$ gives $`a=1.309`$, $`b=0.695`$).
Using that estimate we immediately compute the function $`Q(w)`$ which is plotted in figure 4 for various values of $`ฯต`$.
It is interesting to study the limit $`ฯต\mathrm{}`$ of our formula. Let us suppose that $`y`$ behaves as $`y(q)=aq^\alpha +\mathrm{}.`$ for $`\alpha >1`$ for low $`q`$. In this case $`P(q)`$ will be dominates by the behavior of $`y`$ close to $`q=0`$. We can introduce a cut-off $`\mathrm{\Lambda }`$ such that $`\mathrm{\Lambda }0`$ and $`\mathrm{\Lambda }^{1+\alpha }ฯต\mathrm{}`$.
$$P(q)ฯต\alpha q^{\alpha 1}_q^\mathrm{\Lambda }\mathrm{exp}\left(\frac{ฯต}{\alpha +1}q^{\alpha +1}\right)$$
(35)
which, rescaling the integration variable and sending the cut-off to zero becomes:
$$P(q)=\frac{1}{q}\left(\frac{ฯตaq^{\alpha +1}}{1+\alpha }\right)^{\frac{\alpha }{1+\alpha }}\mathrm{\Gamma }(\frac{1}{1+\alpha },\frac{ฯตa}{\alpha +1}q^{\alpha +1})$$
(36)
where the incomplete gamma function is
$$\mathrm{\Gamma }(n,x)=_x^{\mathrm{}}๐yy^{n1}e^y.$$
(37)
This case is relevant in the spherical models where $`y(q)\frac{a}{4(p3)!}q^{p3}`$ and for the SK model where $`y(q)aq`$.
## V Directed polymers in $`1+1`$ dimension
The natural play ground of the exposed theory are finite dimensional spin glasses. Our analysis predicts a very peculiar dependence of the probability of $`q`$ and $`w`$ on $`ฯต`$. Formulae (28,30) has been derived assuming that the low lying states verify ultrametricity. It is therefore interesting understand what violations of the scaling forms (28,30) can be expected when the ground states structure is nontrivial, but ultrametricity does not hold.
In this paper we study numerically the simple case of directed polymers in random media in 1+1 dimension. The model we use is defined on the square lattice, where the polymer can perform a random walk starting from the origin. On each site of the lattice is defined an passage energy cost which is a Gaussian variable with unit variance, and independent of all the other energies. The properties of this sort of models have been studied extensively , and it is well known that while the low temperature thermodynamics is dominated by a single ground state, there exist many โpure statesโ (i.e. metastable states separated by growing barriers) with typical energy gap with the ground state scaling as $`L^{1/3}`$ (for a polymer of total length $`L`$). The overlap for two polymers of length $`L`$ with a common source in the origin is often defined as the fraction of the monomers passing in the same sites of the lattice in the two polymers. Given the previously mentioned scaling of the energy gap it is natural that the scaling of the coupling in order to have a non trivial $`P(q)`$ as defined in the previous section is
$$ฯต=\eta L^{1/3}.$$
(38)
If we had to suppose the validity of the formula (28) we would conclude that the function $`y(q)`$ for samples of length $`L`$ depends on $`L`$ and scales as $`L^{1/3}`$.
As we stressed, ultrametricity does not hold in this model. Although we did not make a systematic study, we can easily show the lack of ultrametricity generalizing the coupling procedure to a third โreplicaโ, which has a repulsion both with the unperturbed ground state, and with the one obtained with the coupling. For simplicity, we look at the case in which all couplings are equal. In figure 5 we show the overlap between the second and the third replica as a function of the overlap between the first and the third, fixing the overlap between the first and the second to 0.8, and we see no sign of ultrametricity.
In figure 6 we show the function $`Q(w)`$ for various values of $`L`$ and $`ฯต`$ of the form (38). We see that for values of $`L200`$ the expected independence of $`Q(w)`$ of $`L`$ is reasonably obeyed. A close inspection to $`Q(1)`$ however, which represents the probability of $`q=1`$ reveals that this quantity behaves as $`Q(1)=\mathrm{exp}(\eta \chi L^{2/3})`$, with $`\chi =0.85\pm 0.02`$, and that the scaling is violated in proximity of $`w=1`$.
We next try to scale the data with $`ฯต`$ according to the formula (30). In figure 7 we see that the works quite well for small values of $`ฯต`$, while we show in figure 8 that there are important violations for large values of $`ฯต`$.
## VI Conclusions
Monte-Carlo simulations of finite dimensional spin glasses, show a behavior in agreement with RSB . However, the use of Monte-Carlo has been criticized on the ground that one can only equilibrate the system too close to the critical point, where finite size effects are large and could spoil the conclusions about ergodicity breaking in the thermodynamic limit (see however ). It is important therefore to find consequences of RSB at zero temperature. In this paper we have devised some of them.
We have found that a universal formula holds for the probability of the overlap among the uncoupled ground state and the coupled one. The investigation of the validity of that formula in three dimensional systems is not beyond reach with the present technology, and will furnish an important test to understanding the nature of the spin-glass phase of three dimensional systems.
Two main ingredient will be involved in the calculation: the exponential distribution of the states and their ultrametric organization . This implies that the tree of states is described by a single function $`y(q)`$. The function $`y(q)`$ may in principle depend on $`N`$. Mean field theory predicts that $`y(q)`$ remains finite in the thermodynamic limit implying that the energy differences between pure states remain finite in the thermodynamic limit. However, one could envisage systems where both exponential distribution and ultrametricity are valid, but the typical energy differences scale as $`L^\theta `$. In this case, one still have a function $`y(q)`$ which scales as $`L^\theta `$ and in order to measure a nontrivial overlap distribution one needs a coupling of order $`L^\theta `$. The numerical study of the overlap and gap distribution, and the comparison with the formulae found in this paper will give important information about the organization of the states in short range models.
## Acknowledgments
We thank Enzo Marinari, Matteo Palassini and Peter Young for important discussions.
|
warning/0006/quant-ph0006087.html
|
ar5iv
|
text
|
# Conceptual Inadequacy of the Shannon Information in Quantum Measurements
## I Introduction
In classical physics information is represented as a binary sequence, i.e a sequence of bit values, each of which can be either $`1`$ or $`0`$. When we read out information that is carried by a classical system we reveal a certain bit value that exists even before the reading of information is performed. For example, when we read out a bit value encoded as a pit on a compact disk, we reveal a property of the disk existing before the reading process.
This means that in a classical measurement the particular sequence of bit values obtained can be considered to be physically defined by the properties of the classical system measured<sup>*</sup><sup>*</sup>* Even in these cases where classical physics instead of definite measurement results predicts these results with certain probabilities, it is still possible at least in principle, to consider an ensemble of statistically distributed measurement results as revealing corresponding statistically distributed properties of the ensemble of classical systems.. The information read is then measured by the Shannon measure of information which can operationally be defined as the number of binary questions (questions with โyesโ or โnoโ answers only) needed to determine the actual sequence of 0โs and 1โs.
In quantum physics information is represented by a sequence of qubits, each of which is defined in a two-dimensional Hilbert space. If we read out the information carried by the qubit, we have to project the state of the qubit onto the measurement basis $`\{|0,|1\}`$ which will give us a bit value of either 0 or 1. Only in the exceptional case of the qubit in an eigenstate of the measurement apparatus the bit value observed reveals a property already carried by the qubit. Yet in general the value obtained by the measurement has an element of irreducible randomness and therefore cannot be assumed to reveal the bit value or even a hidden property of the system existing before the measurement is performed.
This implies that in a sequence of measurements on qubits in a superposition state $`a|0+b|1`$ $`(|a|,|b|\{0,1\})`$ the particular sequence of bit values 0 and 1 obtained cannot, not even in principleAs theorems like those of Kochen-Specker show, it is fundamentally not possible to assign to a quantum system (noncontextual) properties corresponding to all possible measurements. The theorems assert that for a quantum system described in a Hilbert space of dimension equal to or larger than three, it is possible to find a set of $`n`$ projection operators which represent the yes-no questions about an individual system, such that none of the $`2^n`$ possible sets of answers is compatible with the sum rule of quantum mechanics for orthogonal decomposition of identity (i.e. if the sum of a subset of mutually commuting projection operators is the identity one and only one of the corresponding answers ought to be โyesโ). This means that it is not possible to assign a definite unique answer to every single yes-no question represented by a projection operator independent of which subset of mutually commuting projection operators one might consider it with together. If there are no definite (context-independent) answers to all possible yes-no questions that can be asked about the system then the operational concept of the Shannon measure of information itself, defined as the number of yes-no questions needed to determine the particular answers the system gives, becomes highly problematic., be considered in any way to be defined before the measurements are performed. The non-existence of well-defined bit values prior to and independent of observation suggests that the Shannon measure, as defined by the number of binary questions needed to determine the particular observed sequence 0โs and 1โs, becomes problematic and even untenable in defining our uncertainty as given before the measurements are performed.
Here we will critically analyze the applicability of the axiomatic derivation of the Shannon measure for the case of quantum measurement. We will also show that Shannon information is not useful in defining the information content in a quantum system. In fact we will see that when we try to apply Shannonโs postulate in quantum measurements or when we try to define the information content by the Shannon information a certain element emerges that escapes complete and full description in quantum mechanics. This element is always associated with the objective randomness of individual quantum events and with quantum complementarity. In the end we will briefly discuss a novel and more suitable measure of information . Yet at first we will return to a discussion in more detail of the operational definition of Shannon information to quantum measurements.
## II Discussion of the Operational Definition for a Sequence of Measurements
For classical observations Shannonโs measure of information can conceptually be motivated through an operational approach to the question. We will follow the introduction of Shannonโs measure of information as given by Uffink . Consider an urn filled with N colored balls. There are $`n_1`$, $`n_2`$, โฆ, $`n_m`$ balls with various different colors: black, white, โฆ, red. Now the urn is shaken, and we draw one after the other all balls from the urn. To what extent can we predict the particular color sequence drawn?
Certainly, if all the balls in the urn are of the same color, we can completely predict the color sequence. On the other hand, if the various colors are present in equal proportions and if we have no knowledge about the arrangement of the balls after shaking the urn, we are maximally uncertain about the color sequence drawn. As noticed in one can think of these situations as extreme cases on a varying scale of predictability. For example, for N=4, there is only one color sequence $``$$``$$``$$``$ if all balls are white, 4 possible color sequences $``$$``$$``$$``$, $``$$``$$``$$``$, $``$$``$$``$$``$, $``$$``$$``$$``$, if there are three black and one white ball in the urn, yet 6 possible color sequences $``$$``$$``$$``$, $``$$``$$``$$``$, $``$$``$$``$$``$, $``$$``$$``$$``$, $``$$``$$``$$``$ and $``$$``$$``$$``$ if there are two black and two white balls in the urn. This suggest that the uncertainty we have before drawing about the particular color sequence that will be drawn is defined by the total number of different possible color sequences that are in accordance with the given number of balls with their respective colors in the urn.
Consider now a situation where a long sequence of N balls are drawn from an infinite โseaโ of balls with proportions $`p_1`$, $`p_2`$, โฆ, $`p_m`$ for the different colours in the sea. Then a long sequence contains with high probability about $`p_1N`$ balls of the first colour, $`p_2N`$ balls of the second colour etc. (such a sequence is called typical sequence). The probability to obtain a particular typical sequence (particular colour sequence) is given by
$$p(sequence)=p_1^{p_1N}p_2^{p_2N}\mathrm{}p_m^{p_mN}=\frac{1}{2^{NH}}$$
(1)
where
$$H=\underset{i=1}{\overset{m}{}}p_i\mathrm{log}p_i$$
(2)
is the Shannon information expressed in bits with the logarithm taken to base 2. Consequently, the total number of distinct typical sequences is given by
$$W2^{NH}.$$
(3)
Suppose now that one wishes to identify a specific color sequence of the drawn balls from the complete set of possible color sequences by asking questions to which only โyesโ or โnoโ can be given as an answer. Of course, the number of questions needed will depend on the questioning strategy adopted. In order to make this strategy the most optimal, that is, in order that we can expect to gain maximal information from each yes-or-no question, we evidently have to ask questions whose answers will strike out always half of the possibilities.
Since there are $`W=2^{NH}`$ possible different (typical) color sequences (all of them have equal probability to be drawn), the minimal number of yes-no questions needed is just $`NH`$. Or equivalently, the Shannon information expressed in bits is the minimal number of yes-no questions necessary to determine which particular sequence of outcomes occurs, divided by $`N`$ . A particular color sequence is specified by writing down, in order, the yesโs and noโs encountered in traveling from the root to the specific leaf of the tree as schematically depicted in Fig. 1 for an explicit example with an urn containing black and white balls only.
If instead of balls with pre-assigned colors we consider quantum systems whose individual properties are not defined before the measurements are performed, does the Shannon measure of information still define the information gain in the measurements appropriately? More precisely, we ask here the question whether the total number $`W=2^{NH}`$ of different possible (typical) sequences of outcomes is suitable as a measure of our uncertainty before the sequence of quantum measurements is performed.
In classical physics the behavior of the whole ensemble follows from the behavior of its intrinsic different individual constituents which can be thought of as being defined to any precision. This is not the case in quantum mechanics. The principal indefiniteness, in the sense of fundamental nonexistence of a detailed description of and prediction for the individual quantum event resulting in the particular measurement result, implies that the particular sequence of outcomes specified by writing down, in order, the yesโs and noโs encountered in a row of yes/no questions asked is not defined before the measurements are performed. No definite outcomes exists before measurements are performed and therefore the number of different possible sequence of outcomes does not characterize our uncertainty about the individual system given before measurements are performed.
However, once the sequence of quantum measurements is performed and the measurement results are obtained, the measure of information needed to specify the particular sequence of outcomes realized is defined appropriately by the Shannon measure. In the sense that an individual quantum event manifests itself only in the measurement process and is not precisely defined before measurement is performed, we may speak of โgenerationโ of that specific information in the measurement.
## III Inapplicability of Shannonโs Postulates in Quantum Measurements
As observed by Uffink , an important reason for preferring the Shannon measure of information lies in the fact that it is uniquely characterized by Shannonโs intuitively reasonable postulates. This has been expressed strongly by Jaynes : โOne โฆ important reason for preferring the Shannon measure is that it is the only one that satisfies โฆ \[Shannonโs postulates\]. Therefore one expects that any deduction made from other information measures, if carried far enough, will eventually lead to contradiction.โ A good way to continue our discussion is by reviewing how Shannon, using his postulates, arrived at his famous expression. He writes :
โSuppose we have a set of possible events whose probabilities of occurrence are $`p_1,p_2,\mathrm{},p_n`$. These probabilities are known but that is all we know concerning which event will occur. Can we find a measure of how much โchoiceโ is involved in the selection of the event or how uncertain we are of the outcome?
If there is such a measure, say $`H(p_1,p_2,\mathrm{},p_n)`$, it is reasonable to require of it the following properties:
1. $`H`$ should be continuous in the $`p_i`$.
2. If all the $`p_i`$ are equal, $`p_i=\frac{1}{n}`$, then $`H`$ should be a monotonically increasing function of $`n`$. With equally likely events there is more choice, or uncertainty, when there are more possible events.
3. If a choice be broken down into two successive choices, the original $`H`$ should be the weighted sum of the individual values of $`H`$. The meaning of this is illustrated in Fig. 2. At the left we have three possibilities $`p_1=\frac{1}{2}`$, $`p_2=\frac{1}{3}`$, $`p_3=\frac{1}{6}`$. On the right we first choose between two possibilities each with probability $`\frac{1}{2}`$, and if the second occurs make another choice with probabilities $`\frac{2}{3}`$, $`\frac{1}{3}`$. The final results have the same probabilities as before. We require, in this special case, that
$`H({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{3}},{\displaystyle \frac{1}{6}})=H({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}})+{\displaystyle \frac{1}{2}}H({\displaystyle \frac{2}{3}},{\displaystyle \frac{1}{3}}).`$
The coefficient $`\frac{1}{2}`$ is the weighing factor introduced because this second choice occurs half the time.โ
Shannon then shows that only the function (2) satisfies all three postulates. It is the third postulate which determines the logarithm form of the function and, as we will argue, it is this postulate which leads to problems when quantum measurements are involved.
We now turn to the discussion of Shannonโs postulates. While the first two postulates are natural for every meaningful measure of information, the last postulate might deserve more justification The third Shannon postulate originally formulated as an example was reformulated as an exact rule by Faddeev : For every $`n2`$
$$H(p_1,..,p_{n1},q_1,q_2)=H(p_1,..,p_{n1},p_n)+p_nH(\frac{q_1}{p_n},\frac{q_2}{p_n}),$$
(4)
where $`p_n=q_1+q_2`$.
Without physical interpretation the recursion postulate (the name was suggested in ) (4) is merely a mathematical expression which is certainly necessary for the uniqueness of the function (2) but has no further physical significance. We adopt the following well-known interpretation . Assume the possible outcomes of the experiment to be $`a_1,\mathrm{},a_n`$ and $`H(p_1,\mathrm{},p_n)`$ to represent the amount of information that is gained by the performance of the experiment. Now, decompose event $`a_n`$ into two distinct events $`a_nb_1`$ and $`a_nb_2`$ (โ$``$โ denotes โandโ, thus $`ab`$ denotes a joint event). Denote the probabilities of outcomes $`a_nb_1`$ and $`a_nb_2`$ by $`q_1`$ and $`q_2`$, respectively. Then the left-hand side $`H(p_1,\mathrm{},p_{n1},q_1,q_2)`$ of Eq. (4) represents the amount of information that is gained by the performance of the experiment with outcomes $`a_1,\mathrm{},a_{n1},a_nb_1,a_nb_2`$.
When the outcome $`a_n`$ occurs, the conditional probabilities for $`b_1`$ and $`b_2`$ are $`\frac{q_1}{p_n}`$ and $`\frac{q_2}{p_n}`$ respectively and the amount of information gained by the performance of the conditional experiment is $`H(\frac{q_1}{p_n},\frac{q_2}{p_n})`$. Hence the recursion requirement states that the information gained in the experiment with outcomes $`a_1,\mathrm{},a_{n1},a_nb_1,a_nb_2`$ equals the sum of the information gained in the experiment with outcomes $`a_1,\mathrm{},a_n`$ and the information gained in the conditional experiment with outcomes $`b_1`$ or $`b_2`$, given that the outcome $`a_n`$ occurred with probability $`p_n`$.
This interpretation implies that the third postulate can be rewritten as
$`H`$ $`(p(a_1),\mathrm{},p(a_{n1}),p(a_nb_1),p(a_nb_2))`$ (5)
$`=`$ $`H(p(a_1),\mathrm{},p(a_{n1}),p(a_n))+p(a_n)H(p(b_1|a_n),p(b_2|a_n))`$ (6)
where
$`p`$ $`(a_n)=p(a_nb_1)+p(a_nb_2)`$ (7)
$`p`$ $`(a_nb_1)=p(a_n)p(b_1|a_n)\text{ and}`$ (8)
$`p`$ $`(a_nb_2)=p(a_n)p(b_2|a_n).`$ (9)
Here $`p(b_i|a_n)`$ $`i=1,2`$ denotes the conditional probability for outcome $`a_n`$ given the outcome $`b_i`$ occurred and $`p(a_nb_i)`$ denotes the joint probability that outcome $`a_nb_i`$ occurs.
If we analyze the generalized situation with $`n`$ outcomes $`a_i`$ of the first experiment $`A`$, $`m`$ outcomes $`b_j`$ of the conditional experiment $`B`$ and $`mn`$ outcomes $`a_ib_j`$ of the joint experiment $`AB`$, we may then rewrite the recursion postulate in a short form as
$$H(AB)=H(A)+H(B|A)$$
(10)
where $`H(B|A)=_j^np(a_j)H(b_1|a_j,\mathrm{},b_m|a_j)`$ is the average information gained by observation $`B`$ given that the conditional outcome $`a_j`$ occurred weighted by probability $`p(a_j)`$ for $`a_j`$ to occur.
It is essential to note that the recursion postulate is inevitably related to the manner in which we gain information in a classical measurement. In fact, in classical measurements it is always possible to assign to a system simultaneously attributes corresponding to all possible measurements, here $`a_i`$, $`b_j`$ and $`a_ib_j`$. Also, the interaction between measuring apparatus and classical system can be thought to be made arbitrarily small so that the experimental determination of $`A`$ has no influence on our possibility to predict the outcomes of the possible future experiment $`B`$. In conclusion, the information expected in a classical experiment from the joint experiment $`AB`$ is simply the sum of the information expected from the first experiment $`A`$ and the conditional information of the second experiment $`B`$ with respect to the first, as expressed in Eq. (10).
Therefore, only for the special case of commuting, i.e. simultaneously definite observables, the axiomatic derivation of the Shannon measure of information is applicable and the use of the Shannon information is justified to define the uncertainty given before quantum measurements are performed. However, in general, if $`A`$ and $`B`$ are noncommuting observables, the joint probabilities on the left-hand side of Eq. (5) cannot in principle be assigned to a system simultaneously, and consequently Shannonโs crucial third postulate which is necessary for the uniqueness of Shannonโs measure of information ceases to be well-defined.
Having seen that the third Shannon postulate in general is not applicable in quantum measurements we next introduce two requirements that are immediate consequences of Shannonโs postulates and in which all the probabilities that appear are well-defined in quantum mechanics. We will show that the two requirements are violated by the information gained in quantum measurements implying that the Shannon measure loses its preferential status with respect to alternative expressions when applied to define information gain in quantum measurements.
1. Every new observation reduces our ignorance and increases our knowledge. In his work Shannon offers a list of properties to substantiate that $`H`$ is a reasonable measure of information. He writes: โIt is easily shown that
$`H(AB)H(A)+H(B)`$
with equality only if the events are independent (i.e., $`p(a_ib_j)=p(a_i)p(b_j))`$. The uncertainty of a joint event is less than or equal to the sum of the individual uncertaintiesโ. He continues further in the text: โโฆ we have
$`H(A)+H(B)H(AB)=H(A)+H(B|A).`$
Hence,
$$H(B)H(B|A).$$
(11)
The uncertainty of $`B`$ is never increased by knowledge of $`A`$. It will be decreased unless $`A`$ and $`B`$ are independent events, in which case it is not changedโ (we have changed Shannonโs notation to coincide with that of our work).
2. Information is indifferent on the order of acquisition. The total amount of information gained in successive measurements is independent of the order in which it is acquired, so that the amount of information gained by the observation of $`A`$ followed by the observation of $`B`$ is equivalent to the amount of information gained from the observation of $`B`$ followed by the observation of $`A`$
$$H(A)+H(B|A)=H(B)+H(A|B).$$
(12)
This is an immediate consequence of the recursive postulate which can be obtained when we write the recursion postulate in two different ways depending on whether the observation of $`A`$ is followed by the observation of $`B`$ or vice versa. An explicit example for a sequence of classical measurements is given in Fig. 3.
Are these two requirements satisfied by information gained in quantum measurements? Consider a beam of randomly polarized photons. Filters $`F_{}`$, $`F_{45^{}}`$ and $`F_{}`$ are oriented vertically, at $`+45^{}`$, and horizontally respectively, and can be placed so as to intersect the beam of photons (Fig. 4). If we insert filter $`F_{}`$ the number of photons observed at the detection plate will be approximately half of the number in the incoming beam. The outgoing photons now all have vertical polarization. Notice that the function of filter $`F_{}`$ cannot be explained as a โsieveโ that only lets those photons pass that are already of vertical polarization in the incoming beam. If that were the case, only a certain small number of the randomly polarized incoming photons would have vertical polarization, so we would expect a much larger attenuation of the beam of photons as they pass the filter.
Denote with $`A`$ and $`B`$ properties of the photon to have polarization at $`+45^{}`$ and horizontal polarization, respectively. If $`F_{}`$ is inserted behind the filter $`F_{}`$ we are certain that none of the photons will pass through (Fig. 4a). For a photon with vertical polarization we have complete knowledge of the property $`B`$, i.e. $`H(B)=0`$. Notice that a โsieveโ model could explain this behaviour. If we now insert $`F_{45^{}}`$ between $`F_{}`$ and $`F_{}`$ we observe an effect which cannot be explained by a sieve model where the filter does not change the object. However we now observe a certain number of photons at the detection plate (about $`\frac{1}{4}`$ of the number of photons in the beam passed through $`F_{}`$) as shown in Fig. 4b. In this case our knowledge of the property $`B`$ is not complete anymore.
The acquisition of information about property $`A`$ therefore leads to a decrease of our knowledge about property $`B`$, i.e. $`H(B|A)>0`$. Note that on the photons absorbed by the filter $`F_{45^{}}`$ we cannot measure property $`B`$ subsequently. However already for the subensemble of the photons passing through the filter our uncertainty about property $`B`$ becomes larger than 0 implying $`0=H(B)<H(B|A)`$ which clearly violates requirement (11). Another example of sequence of quantum measurements where requirement (12) is violated is given in Fig. 5. Clearly, violation of the requirements (11) and (12) occurs when the corresponding operators $`A`$ and $`B`$ do not commute.
What is the origin of the violation of the requirements (11) and (12) in quantum measurements? In contrast to a classical measurement which just adds some new knowledge to our knowledge at hand from the previous measurements, in a quantum measurement the gain of the new knowledge is always at the expense of irrecoverable loss of complementary classes of knowledge. This originates from the distinction between โtotalโ and โcompleteโ information in quantum physics. In classical physics the total information about a system is complete. In quantum physics the total information of a system, represented by the state vector, is never complete in the sense that all possible future measurement results are precisely definedYet, we do not hesitate to emphasize that it certainly is complete in the sense that it is not possible to have more information about a system than what can be specified in its quantum state. In fact, the state vector represents that part of our knowledge about the history of a system which is necessary to arrive at the maximum possible set of probabilistic predictions for all possible future observations of the system. For example, a set of complex amplitudes of a $`\psi `$-function is a specific representation of the catalog of our knowledge of the system. This view was assumed by Schrรถdinger who wrote: โSie ((die $`\psi `$-Funktion )) ist jetzt das Instrument zur Vorausage der Wahrscheinlichkeit von Maรzahlen. In ihr ist die jeweils erreichte Summe theoretisch begrรผndeter Zukunfterwartungen verkรถrpert, gleichsam wie in einem Katalog niedergelegt. Translated: โIt (the $`\psi `$-function) is now the means for predicting the probability of measurement results. In it is embodied the momentarily attained sum of theoretically based future expectation, somewhat as laid down in a catalog.โ. In fact, the total information of a quantum system suffices to specify the eigenstate of one nondegenerate (with one-dimensional eigenspaces only) observable only.
For example, the state of a photon passing through filter $`F_{}`$ is specified by the complete knowledge about the property $`A`$ of vertical polarization. If we let a photon in this state pass through filter $`F_{45^{}}`$ as given in Fig. 4b, our knowledge of the photon changes, and therefore its representation, the quantum state, also changes. The total information of a photon in the new state is completely exhausted in specifying property $`B`$ of polarization at $`45^{}`$ and no further information is left to also specify property $`A`$, thus implying unavoidable loss of the previous knowledge about this property. This further implies that the set of future probabilistic predictions specified by the new projected state is indifferent to the knowledge collected from the previous measurements in the whole history of the system. Such a view was assumed by Pauli who writes<sup>ยง</sup><sup>ยง</sup>ยงIn translation: โIn the case of indefiniteness of a property of a system for a certain experimental arrangement (for a certain state of the system) any attempt to measure that property destroys (at least partially) the influence of earlier knowledge of the system on (possibly statistical) statements about later possible measurement results.โ: โBei Unbestimmtheit einer Eigenschaft eines Systems bei einer bestimmten Anordnung (bei einem bestimmten Zustand des Systems) vernichtet jeder Versuch, die betreffende Eigenschaft zu messen, (mindestend teilweise) den Einfluร der frรผheren Kenntnisse vom System auf die (eventuell statistischen) Aussagen รผber spรคtere mรถgliche Messungsergebnisse.โ This clearly makes possible to violate requirements (11) and (12) in quantum measurements.
Here a certain misconception might be put forward that arises from a certain practical point of view. According to that view, for example, complementarity between interference pattern and information about the path of the particle in the double-slit experiment is considered to arise from the fact that any attempt to observe the particle path would be associated with an uncontrollable disturbance of the particle. Such a disturbance in itself would then be the reason for the loss of the interference pattern. In such of view it would be possible to define Shannonโs information for all attributes of the system simultaneously, and the third Shannon postulate, as well as the requirements (11) and (12), would be violated because of the unavoidable disturbance of the system occurring whenever the subsequently measured property $`B`$ is incompatible with the previous one $`A`$. Yet, this is a misconception for two reasons.
Firstly, as theorems like those of Bell or Greenberger, Horne and Zeilinger show, it is not possible, not even in principle, to assign to a quantum system simultaneously observation-independent properties which in order to be in agreement with special relativity have to be local. We therefore cannot speak of a โdisturbanceโ in the measurement process if there are no objective properties to disturb.
Secondly, over the last few years experiments were considered and some already performed, where the reason why no interference pattern arises is not due to any uncontrollable disturbance of the quantum system or the clumsiness of the apparatus. Rather the lack of interference is due to the fact that the quantum state is prepared in such a way as to permit path information to be obtained, in principle, independent of whether the experimenter cares to read it out or not. One line of such research considers the use of micromasers in atomic beam experiments , another one concerns experiments on correlated photon states emerging from nonlinear crystals through the process of parametric-down conversion .
The view that complementarity must be based on the much more fundamentally property of mutual exclusiveness of different classes of information of a quantum system was assumed by Pauli in the analysis of the uncertainty relationsIn translation:โโฆ these relations contain the statement that any precise knowledge of the position of a particle implies a fundamental indefiniteness, not just an unknownness, of the momentum for a consequence and vice versa. The distinction between (fundamental) indefiniteness and unknownness, and the relation between these two notions is decisive for the whole quantum theory.โ: โ โฆ diese Relationen enthalten die Aussage, daร jede genaue Kenntnis des Teilchenortes zugleich eine prinzipielle Unbestimmtheit, nicht nur Unbekanntheit des Impulses zur Folge hat und umgekehrt. Die Unterscheidung zwischen (prinzipieller) Unbestimmtheit und Unbekanntheit und der Zusammenhang beider Begriffe sind fรผr die ganze Quantentheorie entscheidend.โ
## IV Difficulties in Defining the Information Content of a Quantum System
To define the information content of a physical system one might consider different measures of information. However only those measures of information have physical significance according to which the defined information content of the system possesses properties which naturally follow from the physical situation considered. These properties are, for example, invariance under changes of the modes of observation of the system and conservation in time if there is no information exchange with an environment. We show now that the information content of a quantum system, if it is assumed to be measured by the Shannon measure of information, cannot be defined in any way to have these properties.
The classical world appears to be composed of particles and fields, and the properties of each one of these constituents could be specified quite independently of the particular phenomenon discussed or of the experimental procedure a physicist chooses to determine these properties. In other words the properties of constituents of the classical world are noncontextual.
In particular, the total lack of information about a classical pointlike system (with no rotational and internal degrees of freedom) defined as Shannonโs information associated with the probability distribution over the phase space is independent of the specific set of variables chosen to describe the system completely (such as position and momentum, or bijective functions of them) and conserved in time if there is no information exchange with an environment (i.e. if the system is dynamically independent from the environment and not exposed to a measurement)Given the probability distribution $`\rho (\stackrel{}{r},\stackrel{}{p},t)`$ over the phase space the total lack of information of a classical system is defined as
$$H_{total}(t)=d^3\stackrel{}{r}d^3\stackrel{}{p}\rho (\stackrel{}{r},\stackrel{}{p},t)\mathrm{log}\frac{\rho (\stackrel{}{r},\stackrel{}{p},t)}{\mu (\stackrel{}{r},\stackrel{}{p})},$$
(13) where a background measure $`\mu (\stackrel{}{r},\stackrel{}{p})`$ is an additional ingredient that has to be added to the formalism to ensure invariance under change of variables when we consider continuous probability distributions. The conservation of $`H_{total}`$ in time for a system with no information exchange with an environment is implied by the Hamiltonian evolution of a point in phase space.. Operationally the total information content of a classical system can be obtained in the joint measurement of position and momentum, or in successive measurements in which the observation of position is followed by the observation of momentum or vice versa<sup>\**</sup><sup>\**</sup>\**In full analogy with (12) we may write $`H_{total}(\stackrel{}{r},\stackrel{}{p})=H(\stackrel{}{r})+H(\stackrel{}{p}|\stackrel{}{r})=H(\stackrel{}{p})+H(\stackrel{}{r}|\stackrel{}{p})`$..
Contrary to the classical concepts most quantum-mechanical concepts are limited to the description of phenomena within some well-defined experimental context, that is, always restricted to a specific experimental procedure the physicist chooses. In particular the amount of information gained in an individual quantum measurement depends strongly on the specific experimental context. In the optimal experiment when the measurement basis $`|i`$ coincides with the eigenbasis of the density matrix $`\widehat{\rho }`$ of the system: $`\widehat{\rho }|i=w_i|i`$ the amount of information gained is maximal (See for example ). Since in the basis corresponding to the optimal experiment the density operator is represented by a diagonal matrix with elements $`w_i`$, the information gain defined by the Shannon measure equals the von Neumann entropy as given by<sup>โ โ </sup><sup>โ โ </sup>โ โ For a given density matrix $`\widehat{\rho }`$ the von Neumann entropy
$$S(\widehat{\rho })=Tr(\widehat{\rho }\mathrm{log}\widehat{\rho })$$
(14) is widely accepted as a suitable definition for the information content of a quantum system. For a system described in $`N`$-dimensional Hilbert space this ranges from $`\mathrm{log}N`$ for a completely mixed state to 0 for a pure state. The von Neumann entropy has the important property to be invariant under unitary transformations. However, we observe that any function of the form $`Tr(f(\widehat{\rho }))`$ (the operator $`f(\widehat{\rho })`$ is identified by having the same eigenstates as $`\widehat{\rho }`$ and the eigenvalues $`f(w_j)`$, equal to the function values taken at the eigenvalues $`w_j`$ of $`\widehat{\rho }`$.) possesses this invariance property. We also observe that the von Neumann entropy is a property of the quantum state as a whole without explicit reference to information contained in individual measurements.
$$H=\underset{i}{}w_i\mathrm{log}w_i=Tr(\widehat{\rho }\mathrm{log}\widehat{\rho }).$$
(15)
This has the important property to be invariant under unitary transformations $`\widehat{\rho }\widehat{U}\widehat{\rho }\widehat{U}^+`$. The invariance under unitary transformations implies invariance under the change of the representation (basis) of $`\widehat{\rho }`$ and conservation in time if there is no information exchange with an environment. The later precisely means that if we perform the optimal experiments both at time $`t_0`$ and at some future time $`t`$, Shannonโs information measures associated to the optimal experiments at the two times will be the same, i.e.
$$H(t)=\underset{i}{}w_i(t)\mathrm{log}w_i(t)=\underset{i}{}w_i\mathrm{log}w_i=H(t_0).$$
(16)
Here, the eigenvalues of the density matrix at time $`t`$ are $`w_i(t)=w_i`$.
However, without the additional knowledge of the eigenbasis of the density matrix $`\widehat{\rho }`$ we cannot find the optimal experiment and obtain directly the Shannon information associated. Also, all the statistical predictions that can be made for the optimal measurement are the same as if we had an ordinary (classical) mixture, with fractions $`w_i`$ of the systems giving with certainty results that are associated to the eigenvectors $`|i`$. In this sense the optimal measurement is a classical type measurement and therefore in this particular case, and only then, Shannonโs measure defines the information gain in a measurement appropriately<sup>โกโก</sup><sup>โกโก</sup>โกโกConsider a situation where instead using of single systems to send information to the receiver a sender uses a sequence of $`N`$ systems where each individual system is drawn from an ensemble of pure states $`\{|\psi _1,\mathrm{},|\psi _n\}`$, with frequency of occurrence $`\{w_1,\mathrm{},w_n\}`$ respectively. It was shown in that for sufficiently large $`N`$ there are $`2^{NS(\widehat{\rho })}`$ highly distinguishable sequences of pure states which become mutually orthogonal as $`N\mathrm{}`$. Here $`S(\widehat{\rho })=Tr(\widehat{\rho }\mathrm{log}\widehat{\rho })`$ is the von Neumann entropy and $`\widehat{\rho }=_i^nw_i|\psi _i\psi _i|`$. This means that if the sender uses a sequence consisting of a choice of states that respects the a priori frequencies $`w_i`$, and the receiver distinguishes whole sequences rather than individual states, then the (Shannon) information transmitted per system can be made arbitrary close to $`S(\widehat{\rho })`$. Here again the total density matrix $`\widehat{\rho }^N`$ of $`N`$ systems can be made arbitrary close to the one as if we had a classical mixture of the $`2^{NS(\widehat{\rho })}`$ sequences of states.. Considering also our previous discussion it is therefore not surprising that Shannonโs measure is useful only when applied to measurements which can be understood as classical measurements.
Which set of individual measurements should we perform and how to combine individual measures of information obtained in the set in order to arrive at the information content of a quantum system if we do not know the eigenbasis of the density matrix? Quantum complementarity implies that the total information content of the system might be partially encoded in different mutually exclusive (complementary) observables. These have the property that complete knowledge of the eigenvalue of any one of the observables excludes any knowledge about the eigenvalues of all other observables. Such a set of observables for a spin-1/2 particle can for example be spin components along orthogonal directions.
We consider now a quantum system described in $`n`$ dimensional Hilbert space and we denote a complete set of $`m`$ mutually complementary observables<sup>\**</sup><sup>\**</sup>\**To specify a system described by a $`n\times n`$ density matrix completely one needs $`n^21`$ independent real numbers. Any individual, complete measurement (we consider here only complete measurements, i.e., where operators associated to the measurements are without degeneracy) with $`n`$ possible outcomes defines $`n1`$ independent probability values (the sum of all probabilities for all possible outcomes in an individual experiment is one). Therefore, just on the basis of counting the number of independent variables, we expect that the number of different measurements we need in order to determine the density matrix completely is $`\frac{n^21}{n1}=n+1`$. Ivanovic , and Wootters and Fields demonstrated the existence of exactly $`n+1`$ mutually complementary observables by an explicit construction in the cases of $`n`$ prime and $`n=2^k`$. by $`\{\widehat{A},\widehat{B},\mathrm{}\}`$. The property of mutual expansiveness implies that if the system is in an eigenstate of one of the observables, for example, in the eigenstate $`|a_j`$ of the observable $`\widehat{A}`$ and we measure any other observable from the set, say $`\widehat{B}`$, projecting the system onto states $`\{|b_1,\mathrm{},|b_i,\mathrm{},|b_n\}`$, the individual outcome is completely random (all measurement results are equally probable)
$$|a_j|b_i|^2=\frac{1}{n}i,j.$$
(17)
It was shown in that the density matrix of the system can fully be reconstructed if one performs a complete set of mutually complementary observations. This suggest that the total information content of a quantum system represented by a density matrix $`\widehat{\rho }`$ is all obtainable from a complete set of mutually complementary measurements. To obtain the total information one however cannot perform the set of measurements successively because, unlike the classical case, the information obtained in successive quantum measurements depends on the order of its acquisition (see Fig. 5 and discussion above). Instead it seems that any attempt to obtain the total information content of a quantum system has to be related to the complete set of mutually complementary experiments performed on systems that are all in the same quantum state.
We suggest that it is therefore natural to require that the total information content in a system in the case of quantum systems is sum of the individual amounts of information over a complete set of $`m`$ mutually complementary observables. As already mentioned above, for a spin-1/2 particle these are three spin projections along orthogonal directions. If we define the information gain in an individual measurement by the Shannon measure the total information encoded in the three spin components is given by
$$H_{total}:=H_1(p_x^+,p_x^{})+H_2(p_y^+,p_y^{})+H_3(p_z^+,p_z^{}).$$
(18)
Here, e.g. $`p_x^+`$ is the probabilities to find the particle with spin up along direction $`x`$.
Considering now an explicit example we will show that the total information $`H_{total}`$ based on the Shannon measure is in general not invariant under unitary transformations. We calculate (18) for a spin-1/2 particle in the state $`|\psi =\mathrm{cos}\theta /2|z++\mathrm{sin}\theta /2|z`$ and we find that
$`H`$ $`{}_{total}{}^{}=`$ (19)
$``$ $`{\displaystyle \frac{1\mathrm{sin}\theta }{2}}\mathrm{log}{\displaystyle \frac{1\mathrm{sin}\theta }{2}}{\displaystyle \frac{1+\mathrm{sin}\theta }{2}}\mathrm{log}{\displaystyle \frac{1+\mathrm{sin}\theta }{2}}`$ (20)
$``$ $`\mathrm{cos}^2{\displaystyle \frac{\theta }{2}}\mathrm{log}\left(\mathrm{cos}^2{\displaystyle \frac{\theta }{2}}\right)\mathrm{sin}^2{\displaystyle \frac{\theta }{2}}\mathrm{log}\left(\mathrm{sin}^2{\displaystyle \frac{\theta }{2}}\right)+1`$ (21)
depends on the parameter $`\theta `$, thus being not invariant under unitary transformations. This associates a number of highly counter-intuitive properties to $`H_{total}`$: 1) it can be different for states of the same purity (e.g. it takes its maximal value of 2 bits of information for $`\theta =0`$ and it takes its minimal value of 1.36 bits for $`\theta =\pi /4`$); 2) it changes in time even for a system completely isolated from the environment where no information can be exchanged with environment; 3) it can take different values for different sets of the three orthogonal spin projections. These unnatural properties we see again as a strong indications for inadequacy of the Shannon measure to define the information gain in an individual quantum measurement.
## V A Suggested Alternative Measure of Information
We suggest that it is natural to require that the information content of the quantum system defined as a sum of individual measures over a complete set of mutually complementary measurements is invariant under unitary transformations. Having shown that this cannot be achieved with the Shannon measure of information we now introduce a new measure of information that differs both mathematically and conceptually from Shannonโs measure of information and according to which the information content has invariance property.
The new measure of information is a quadratic function of probabilities<sup>\*โ </sup><sup>\*โ </sup>\*โ Expressions of the general type of Eq. (23) were studied in detail by Hardy, Littlewood and Pรณlya . They introduced a general class of mathematical expressions
$$M_\alpha =\left(\underset{i=1}{\overset{n}{}}p_i^\alpha \right)^{\alpha 1}\text{ for }0\alpha \mathrm{}$$
(22) that from the point of view of information theory all can be assumed to quantify information properly. These expressions are also closely related to Tsallisโs nonextensive entropy $`S_\alpha =\frac{1}{1\alpha }_{i=1}^n(p_i^\alpha 1)`$ and Rรกnyiโs entropy $`H_\alpha =\frac{1}{1\alpha }\mathrm{log}_{i=1}^np_i^\alpha .`$
$$I(p_1,\mathrm{},p_n)=\underset{i=1}{\overset{n}{}}\left(p_i\frac{1}{n}\right)^2,$$
(23)
and it takes into account that for quantum systems the only features known before an experiment is performed are the probabilities for various events to occur (See for discussion; there a specific normalization factor in expression (23) was used resulting in maximally $`k`$ bits for $`n=2^k`$ possible outcomes). The measure $`I(p_1,\mathrm{},p_n)`$ takes its maximal value of $`(n1)/n`$ if one $`p_i=1`$ and it takes its minimal value of 0 when all $`p_i`$ are equal.
The important property of the new measure of information is that the total information defined with respect to it is invariant under unitary transformations. Using Eq. (17) one obtains that the sum over individual measures of information of mutually complementary observations results in
$`I_{total}`$ $`:=`$ $`{\displaystyle \underset{j=1}{\overset{m}{}}}I(p_1^j,\mathrm{},p_n^j)`$ (24)
$`=`$ $`{\displaystyle \underset{j=1}{\overset{m}{}}}{\displaystyle \underset{i=1}{\overset{n}{}}}\left(p_i^j{\displaystyle \frac{1}{n}}\right)^2=Tr\widehat{\rho }^2{\displaystyle \frac{1}{n}},`$ (25)
for a system described by the density matrix $`\widehat{\rho }`$. Here $`p_i^j`$ denotes the probability to observe the $`i`$-th outcome of the $`j`$-th observable. The total information content of the system therefore might all be encoded in one single observable or, alternatively it might be partially encoded in all $`m`$ mutually complementary observables. For a composite system in a product state the total information can all be encoded in individual systems constituting the composite system or, alternatively in the extreme case of maximally entangled states it can all be encoded in joint properties of the systems with no information left in individual systems .
Independent of the various possibilities to encode information the total information content of the system cannot fundamentally exceed the maximal possible amount of information that can be encoded in an individual observable $`[=(n1)/n]`$. This upper limit is reached when the system is in the pure state. When the system is in a completely mixed state the total information takes its minimal value of $`0`$.
The property of invariance under unitary transformations implies that the total information content of the system does not dependent of the particular set of mutually complementary observables; it is a characterization of the state of the system alone, not of the specific reference set of complementary observables. Furthermore, since evolution in time is described by a unitary operation the total information of the system is conserved in time if there is no information exchange with the environment.
We would like to note that the total information (24) was used in to study the transfer of entanglement and information for quantum teleportation of an unknown entangled state through noisy quantum channels. The total information (24) belongs to the set of quantum counterparts of nonextensive entropies finding its application in increasing number of problems in quantum physics, e.g. description and controlling of laser cooling , a non-extensive approach to the decoherence problem , description and quantifying of entanglement, and deducing criterions for separability of density matrices .
## Conclusions
In this work we have stressed some conceptual difficulties arising when Shannonโs notion of information is applied to define information gain in a quantum measurement. In particular we find that the axiomatic derivation of Shannonโs measure of information is not applicable in quantum measurements in general. We also show that the information content of a quantum system defined according to Shannonโs measure possesses some strongly non-physical properties. We argue that these difficulties in defining the information gain in quantum measurement by the Shannon measure of information arise whenever it is not possible, not even in principle, to assume that attributes observed are assigned to the quantum system before the observation is performed.
Having critized Shannonโs measure of information as being not appropriate for identifying the information gain in quantum measurement we proposed a new measure of information in quantum mechanics that both mathematically and conceptually differs from Shannonโs measure of information. While Shannonโs information is applicable when measurement reveals a preexisting property, the new measure of information takes into account that for quantum systems the only features known before an experiment is performed are the probabilities for various events to occur. In general, which specific event occurs is objectively random.
The total information content of a quantum system defined according to the new measure of information as the sum of the individual measures of information for mutually complementary observations is invariant under unitary transformations. This implies that the total information content of the system is invariant under transformation from one complete set of complementary variables to another and is conserved in time if there is no information exchange with an environment.
## Acknowledgment
In the previous version we did not make proper full reference to the work of J. Uffink . We would like to thank J. Uffink for pointing out this inadequancy as well as an error in our previous Eq. (1). We also thank C. Simon for helpful comments and discussions. This work has been supported by the Austrian Science Foundation FWF, Project No. F1506 and the US National Science Foundation NSF Grant No. PHY 97-22614.
|
warning/0006/nucl-th0006008.html
|
ar5iv
|
text
|
# Seniority isomerism in proton-rich ๐=82 isotones and its indication to stiffness of the ๐=64 core
## I Introduction
Through recent experiments on unstable nuclei, it has been recognized that the nuclear magic numbers are not rigorous and somewhat depend on $`Z`$ and $`N`$ . The magicity observed around the $`\beta `$-stable line may disappear in a region far from the stability. For instance, the magicity of $`N=8`$ no longer holds in the neutron-rich nucleus <sup>11</sup>Be. Although there has been no clear evidence, it is also of interest whether new magic numbers emerge in proton- or neutron-rich region. So-called submagic numbers such as $`Z=40`$ and $`Z=64`$ have been known, which have been distinguished from the magic numbers partly because their magicity disappears as $`Z`$ or $`N`$ changes. However, we now know that even the usual magic numbers depend more or less on $`Z`$ or $`N`$. A question should be recast: what is the difference between magic numbers and submagic numbers? In this respect, it is worthwhile reinvestigating the stiffness of the subshell closure.
The <sup>146</sup>Gd nucleus shows several indications of the $`Z=64`$ subshell closure (e.g. relatively high excitation energy of $`2_1^+`$. In the $`Z>64`$, $`N=82`$ isotones, high-spin isomers with $`J^\pi =10^+`$ (for even-$`Z`$ nuclei) and $`27/2^{}`$ (for odd-$`Z`$ nuclei) have systematically been observed . In connection to these isomers, the single-$`j`$ shell model with the $`\pi 0h_{11/2}`$ orbit was successfully applied to the $`Z>64`$, $`N=82`$ isotones . In the single-$`j`$ shell model, the seniority reduction formula (SRF) is available for the $`E2`$ decay strengths of the high-spin isomers. The SRF had predicted strong hindrance for the decay strengths of the isomers around $`Z=70`$, which is in coincidence with the measured $`E2`$ properties of the $`10^+`$ and $`27/2^{}`$ isomers. At a glance, this seems to indicate that the $`Z=64`$ subshell is stiff enough for <sup>146</sup>Gd to be treated as an inert core. On the other side, the stiffness of the $`Z=64`$ core has been argued so far. For instance, by analyzing the excitation energy of the $`10^+`$ state in <sup>146</sup>Gd as well as those in the $`Z>64`$ isotones, significant pair excitation across $`Z=64`$ was insisted .
In this article, we shall investigate the $`10^+`$ and $`27/2^{}`$ isomers in the $`Z>64`$, $`N=82`$ nuclei, primarily focusing on the stiffness of the $`Z=64`$ core. For the decay strengths of the isomers, we extend the SRF so that it could apply to the multi-$`j`$ cases. This formula shows that the decay strengths reflect the stiffness of the $`Z=64`$ core. If the approximate degeneracy among the $`0h_{11/2}`$, $`2s_{1/2}`$ and $`1d_{3/2}`$ orbits is taken into consideration, the hindrance of the $`E2`$ strengths of the isomers turns out to indicate the presence of the pair excitation across $`Z=64`$.
## II Single-$`j`$ shell model for $`Z>64`$, $`N=82`$ isotones
The proton-rich $`N=82`$ isotones have been explored experimentally. After the discovery of the $`Z=64`$ submagic nature at <sup>146</sup>Gd , several low-lying levels have been established up to <sup>154</sup>Hf . In this region, the excitation energies of the yrast states are nearly constant from nucleus to nucleus, both for even-$`Z`$ (<sup>148</sup>Dy, <sup>150</sup>Er, <sup>152</sup>Yb and <sup>154</sup>Hf) and odd-$`Z`$ (<sup>149</sup>Ho, <sup>151</sup>Tm and <sup>153</sup>Lu) isotones. Furthermore, high-spin isomers were observed systematically; $`10^+`$ isomers for the even-$`Z`$ isotones around $`E_x3`$ MeV, and $`27/2^{}`$ isomers for the odd-$`Z`$ isotones around $`E_x2.5`$ MeV.
Whereas state-of-the-art shell model calculations with a realistic effective interaction have been applied to the $`Z64`$, $`N=82`$ isotones , there have not been many theoretical studies in the $`Z>64`$ region. Lawson carried out a single-$`j`$ shell model calculation with $`\pi 0h_{11/2}`$ on top of the <sup>146</sup>Gd core . The residual interaction was empirically determined from the experimental energy levels of <sup>148</sup>Dy. The levels of the $`Z66`$ isotones were reproduced to a certain extent, apart from the odd-parity levels for the even-$`Z`$ nuclei and the even-parity ones for the odd-$`Z`$ nuclei, which are outside the model space. It is noted that, while the measured excitation energies of the $`2^+`$, $`8^+`$ and $`10^+`$ states gradually decrease as $`Z`$ increases, this tendency is not reproduced in the single-$`j`$ model.
In the single-closed nuclei, it has been known that the seniority $`v`$ is conserved to a good approximation. This is true also in Lawsonโs results. The $`10^+`$ and $`27/2^{}`$ isomers decay via the $`E2`$ transition. The $`10^+`$ isomers and their daughters $`8^+`$ have the seniority $`v=2`$, which is carried by the $`(0h_{11/2})^2`$ configuration. Similarly, the $`27/2^{}`$ isomers and their daughters $`23/2^{}`$ have $`v=3`$. The $`E2`$ transition is usually described by a one-body operator. The seniority reduction formula (SRF) is well-known in the single-$`j`$ configuration. By representing the $`0h_{11/2}`$ orbit by $`\xi `$, the SRF for the seniority-conserving $`E2`$ transitions gives
$$\xi ^nvJ_f^\pi T(E2)\xi ^nvJ_i^\pi =\frac{\mathrm{\Omega }_\xi n}{\mathrm{\Omega }_\xi v}\xi ^vvJ_f^\pi T(E2)\xi ^vvJ_i^\pi ,$$
(1)
where $`\mathrm{\Omega }_\xi =j_\xi +1/2`$. In the present case $`j_\xi =11/2`$ and $`\mathrm{\Omega }_\xi =6`$. In Lawsonโs model the particle number $`n`$ should be $`Z64`$. Equation (1) shows hindrance of the $`E2`$ strengths by the factor $`[(6n)/(6v)]^2`$ when $`n`$ deviates from $`v`$. This hindrance factor gives parabola behavior of $`B(E2)`$ as a function of $`Z`$ and leads to a remarkably long lifetime around $`Z=70`$, i.e. <sup>152</sup>Yb. This stabilization mechanism is called seniority isomerism.
The experimental data on the $`B(E2)`$ values of the isomers well fit to the parabola in the $`66Z70`$ region. Furthermore, both $`E2`$ strengths of the $`10^+`$ and the $`27/2^{}`$ isomers are described by a single effective charge ($`e_{\mathrm{eff}}1.5e`$). In particular, the strong hindrance of the $`E2`$ transition is actually detected for <sup>152</sup>Yb, with $`B(E2;10^+8^+)=0.9\pm 0.1e^2\mathrm{fm}^4`$ . In comparison with the data, the $`E2`$ strengths are overestimated for <sup>153</sup>Lu and <sup>154</sup>Hf in the single-$`j`$ model to a certain extent, as will be shown later. This discrepancy should be attributed to an effect of the orbits other than $`0h_{11/2}`$.
## III Extension of the seniority reduction formula
Despite its success in predicting the seniority isomerism, the single-$`j`$ model will be too simple to be realistic, since the $`\pi 0h_{11/2}`$ orbit is not isolated. In the odd-$`Z`$ $`N=82`$ isotones, $`1/2^+`$ and $`3/2^+`$ states are present in the vicinity of the $`11/2^{}`$ states, indicating the approximate degeneracy among the proton orbits $`0h_{11/2}`$, $`2s_{1/2}`$ and $`1d_{3/2}`$. A certain number of levels with opposite parities (odd-parity levels for the even-$`Z`$ isotones and even-parity ones for the odd-$`Z`$ isotones) are also observed in the low energy regime, which cannot be described in the $`\pi 0h_{11/2}`$ single-$`j`$ model. We shall reinvestigate the $`10^+`$ and $`27/2^{}`$ isomers in the multi-$`j`$ shell model framework.
There is no evidence for a breakdown of the neutron magic number $`N=82`$ in the low energy region, except for a few states relevant to the octupole collectivity. We hereafter maintain the $`N=82`$ inert core. For the proton degrees-of-freedom, the $`50<Z<82`$ major shell is considered at largest.
While the seniority isomerism in this region has been discussed based on the SRF for the single-$`j`$ orbit $`\pi 0h_{11/2}`$, in the following we show that the formula (1) can be extended to the multi-$`j`$ model space with a simple modification.
Let us define the seniority in the multi-$`j`$ space by the sum of the seniorities of each orbit, $`v=_jv_j`$. The seniority is expected to be a good quantum number in single-closed nuclei, at least for their low-lying states. In the high-spin isomers under interest, the seniority is carried only by the $`\pi 0h_{11/2}`$ orbit, to a good approximation. In the $`50<Z<82`$ major shell, $`J^\pi =10^+`$ with $`v=2`$ is uniquely formed by the $`(0h_{11/2})^2`$ configuration, and $`J^\pi =27/2^{}`$ with $`v=3`$ by $`(0h_{11/2})^3`$. The decays of the isomers occur via the $`E2`$ transition without changing the seniority. Within this major shell, the $`8^+`$ state with $`v=2`$, the final state of the $`10^+`$ decay, also has the $`(0h_{11/2})^2`$ configuration. The $`23/2^{}`$ state having $`v=3`$, the daughter of the $`27/2^{}`$ decay, is predominantly $`(0h_{11/2})^3`$. Although this state may have an admixture of $`(0g_{7/2})^2(0h_{11/2})^1`$ and $`(0g_{7/2})^1(1d_{5/2})^1(0h_{11/2})^1`$, the admixture will be small, because these configurations need excitation across $`Z=64`$ by two protons. Moreover, the remaining part consisting of $`0^+`$ pairs is expected to have almost identical structure between the isomers and their daughter states. This is in accordance with the spherical BCS or Talmiโs generalized-seniority picture , where quasiparticles are defined on top of the coherent $`0^+`$ pairs distributing over the valence orbits. Keeping this situation in mind, we derive an extended formula in somewhat general manner.
Suppose that (a) the seniority is a good quantum number, (b) for a seniority-conserving $`E2`$ transition, the seniority is carried by a single orbit (labelled by $`\xi `$) both for the initial and the final states of the transition, and (c) the wave functions of the paired particles are identical between the two states. The condition (c) will be given below in more definitive manner. We represent all valence orbits other than $`\xi `$ by $`r`$. In the present $`N=82`$ case, $`\xi =\pi 0h_{11/2}`$ and $`r=\pi (0g_{7/2}1d_{5/2}1d_{3/2}2s_{1/2})`$. The shell model bases are decomposed into the product of the $`\xi ^{n_\xi }`$ and $`r^{n_r}`$ configurations, where the valence particle number is given by $`n=n_\xi +n_r`$. Because of (b), the seniorities of the $`\xi `$ and $`r`$ subspaces are $`v_\xi =v`$ and $`v_r=0`$, respectively. The initial and final states are expanded as
$$|(\xi r)^nvJ_i^\pi =\underset{n_\xi (v),\alpha }{}c_{n_\xi \alpha }|\xi ^{n_\xi }v_\xi =vJ_i^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+,$$
(2)
and
$$|(\xi r)^nvJ_f^\pi =\underset{n_\xi (v),\alpha }{}c_{n_\xi \alpha }|\xi ^{n_\xi }v_\xi =vJ_f^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+.$$
(3)
Here $`\alpha `$ represents composition of the $`0^+`$ pairs within the $`r^{nn_\xi }`$ configuration. For instance, $`\alpha `$ distinguishes $`(0g_{7/2}1d_{5/2})^{14}(1d_{3/2})^2(2s_{1/2})^2`$ from $`(0g_{7/2}1d_{5/2})^{14}(1d_{3/2})^4`$. The condition (c) is defined as the expansion coefficients ($`c_{n_\xi \alpha }`$) are equal between $`|J_i^\pi `$ and $`|J_f^\pi `$. For zero-range interactions like the SDI, this condition results from (a) and (b), as is verified in Appendix. The normalization yields
$$\underset{n_\xi ,\alpha }{}c_{n_\xi \alpha }^2=1.$$
(4)
Because of the condition (c), the occupation number on the orbit $`\xi `$ is equal between the initial and the final states:
$$N_\xi =\underset{n_\xi ,\alpha }{}c_{n_\xi \alpha }^2n_\xi ,$$
(5)
where $`N_\xi `$ stands for the number operator on $`\xi `$.
Since the $`r`$ subspace carries no seniority under the condition (b), the $`E2`$ transition is forbidden within this subspace. Namely, in the seniority-conserving $`E2`$ transition, the $`r`$ subspace behaves as a spectator. The $`E2`$ matrix element is then written as
$$(\xi r)^nvJ_f^\pi T(E2)(\xi r)^nvJ_i^\pi =\underset{n_\xi ,\alpha }{}c_{n_\xi \alpha }^2\xi ^{n_\xi }v_\xi =vJ_f^\pi T(E2)\xi ^{n_\xi }v_\xi =vJ_i^\pi .$$
(6)
This $`E2`$ transition is a non-collective one, contributed only by the $`\xi `$ orbit. Substitution of the SRF for the orbit $`\xi `$ (see Eq. (1)) into the right-hand side (RHS) yields
$$(\xi r)^nvJ_f^\pi T(E2)(\xi r)^nvJ_i^\pi =\underset{n_\xi ,\alpha }{}c_{n_\xi \alpha }^2\frac{\mathrm{\Omega }_\xi n_\xi }{\mathrm{\Omega }_\xi v}\xi ^vvJ_f^\pi T(E2)\xi ^vvJ_i^\pi ,$$
(7)
with $`\mathrm{\Omega }_\xi j_\xi +1/2`$. Because of Eqs. (4) and (5), we finally obtain
$$(\xi r)^nvJ_f^\pi T(E2)(\xi r)^nvJ_i^\pi =\frac{\mathrm{\Omega }_\xi N_\xi }{\mathrm{\Omega }_\xi v}\xi ^vvJ_f^\pi T(E2)\xi ^vvJ_i^\pi .$$
(8)
Equation (8) links the $`E2`$ matrix element to $`N_\xi `$, occupation number on the orbit $`\xi `$. If the effective charge parameter in $`T(E2)`$ is fixed in advance, the $`E2`$ matrix element is determined only from $`N_\xi `$. Conversely, $`N_\xi `$ can be extracted from the $`E2`$ matrix element. What determines $`N_\xi `$ is $`c_{n_\xi \alpha }`$, which represents the configuration mixing due to the pairing correlation. Thus the $`E2`$ strengths of the isomers are a pairing property, sensitive to the mixing via the pairing interaction.
Compare the formula (8) to the SRF for the single-$`j`$ orbit (1). Although the multi-$`j`$ matrix element is under discussion, the only difference in the RHS is that the particle number $`n`$ is replaced by the expectation value $`N_\xi `$. We shall call Eq. (8) extended seniority reduction formula (ExSRF). The hindrance of the transition strength occurs via the factor $`[(\mathrm{\Omega }_\xi N_\xi )/(\mathrm{\Omega }_\xi v)]^2`$, in parallel to the argument in the single-$`j`$ case, and extraordinarily long lifetime is expected if $`N_\xi \mathrm{\Omega }_\xi `$. The ExSRF (8) obviously contains the single-$`j`$ formula (1) as a limiting case. Equation (8) reduces to Eq. (1) if $`c_{n_\xi \alpha }=1`$ for $`n_\xi =n`$ and $`0`$ for the others. Still the difference from the single-$`j`$ case should be remarked. Even when the seniority is conserved, there could be configuration mixing due to the pairing correlations. While the SRF (1) in the single-$`j`$ model requires that any mixing should be negligible, the ExSRF (8) holds with the pairing mixing. The present conditions to the $`E2`$ hindrance are thereby much more realistic than in the single-$`j`$ case, and the hindrance due to $`N_\xi `$ may be found in a variety of the single-closed nuclei and their neighbors. We shall call this mechanism extended seniority isomerism.
Blomqvist suggested, without proof, that $`n`$ in the SRF (1) can be reinterpreted as the occupation number . Discussion based on the BCS approximation was given in Ref. . The BCS argument leads to the factor $`(u_\xi ^2v_\xi ^2)`$ in terms of the $`u`$\- and $`v`$-coefficients , which is proportional to $`(\mathrm{\Omega }_\xi N_\xi )`$. However, the degree of the approximation was not clear enough. The BCS approximation presumes coherent pairing and ignores some dependence on the seniority (e.g. the seniority-dependence in the denominator of Eq. (8)). On the other hand, we have derived the ExSRF in more rigorous and general manner, which is exact as far as the conditions (aโc) are satisfied.
We here comment on the relation of the ExSRF (8) to the multi-$`j`$ quasi-spin (QS) formula for the degenerate single-particle orbits . The multi-$`j`$ QS formula is available when the pair distributes over all the valence orbits with equal amplitudes. We then have
$$\frac{n_\xi v}{nv}=\frac{\mathrm{\Omega }_\xi v}{\mathrm{\Omega }v},$$
(9)
where we use the notation
$$\mathrm{\Omega }=\underset{j(\xi ,r)}{}\mathrm{\Omega }_j=\underset{j(\xi ,r)}{}(j+1/2).$$
(10)
By employing Eqs. (4) and (9), Eq. (7) reduces to
$$(\xi r)^nvJ_f^\pi T(E2)(\xi r)^nvJ_i^\pi =\frac{\mathrm{\Omega }n}{\mathrm{\Omega }v}\xi ^vvJ_f^\pi T(E2)\xi ^vvJ_i^\pi .$$
(11)
Because of the condition (b), Eq. (11) is equivalent to the multi-$`j`$ QS formula
$$(\xi r)^nvJ_f^\pi T(E2)(\xi r)^nvJ_i^\pi =\frac{\mathrm{\Omega }n}{\mathrm{\Omega }v}(\xi r)^vvJ_f^\pi T(E2)(\xi r)^vvJ_i^\pi .$$
(12)
We now return to the case of the $`N=82`$ isotones. Since $`\xi =0h_{11/2}`$ (thereby $`\mathrm{\Omega }_\xi =6`$), the $`10^+`$ or $`27/2^{}`$ isomer has remarkably long life for a nucleus satisfying $`N_{0h_{11/2}}6`$. Namely, the observed long lifetime of the $`10^+`$ isomer in <sup>152</sup>Yb implies $`N_{0h_{11/2}}6`$. As far as $`2s_{1/2}`$ and $`1d_{3/2}`$ lie closely to $`0h_{11/2}`$, there should be mixing among these orbits due to the pairing interaction, causing decrease of $`N_{0h_{11/2}}`$. However, it can be compensated by the excitation from $`0g_{7/2}`$ or $`1d_{5/2}`$ to $`0h_{11/2}`$, which increases $`N_{0h_{11/2}}`$. As we shall discuss in the following sections, this should be what happens in the isomers in the $`Z>64`$, $`N=82`$ isotones.
## IV Multi-$`j`$ shell model for $`Z64`$, $`N=82`$ isotones
### A Model space
In this section, we present how the properties of the high-spin isomers are described, by a calculation in the multi-$`j`$ shell model framework. As discussed in the preceding section, the model space should include all the five orbits in the $`50<Z<82`$ major shell. Large-scale shell model calculations were carried out for the $`62Z65`$, $`N=82`$ isotones with moderate truncation , as well as for the $`Z64`$, $`N=82`$ isotones in the full major shell . On the other hand, our main purpose is to illustrate the extended seniority isomerism in the $`Z>64`$, $`N=82`$ isotones. In order to avoid time-consuming computations, we adopt relatively small space by truncation.
The space for diagonalization is truncated as follows. Partially maintaining the $`Z=64`$ subshell structure, we restrict the excitation out of the $`0g_{7/2}`$ and $`1d_{5/2}`$ orbits to four particles. Furthermore, the total seniority is limited to $`v3`$ ($`v2`$) for the odd-$`Z`$ (even-$`Z`$) nuclei. The seniorities of the $`10^+`$ and the $`27/2^{}`$ states are pure in this space as well as those of their decay daughters; the condition (a) in Section III is satisfied. The condition (b) is exact for the $`10^+`$ decay, while the final state $`23/2^{}`$ of the $`27/2^{}`$ decay has a small admixture of the $`(0g_{7/2})^2(0h_{11/2})^1`$ and $`(0g_{7/2})^1(1d_{5/2})^1(0h_{11/2})^1`$ configurations, as stated in Section III.
### B Energy levels
The shell model Hamiltonian is written as
$$H=E_0+\underset{j}{}ฯต_jN_j+V.$$
(13)
Here $`E_0`$ is a constant shifting the origin of the energy, $`ฯต_j`$ represents the single-particle energy of the orbit $`j`$, and $`N_j`$ the number operator on $`j`$. The residual two-body interaction is denoted by $`V`$, for which we adopt the modified surface-delta interaction (MSDI),
$`V=4\pi A_{T=1}{\displaystyle \underset{\lambda }{}}Y^{(\lambda )}(\widehat{๐ซ}_1)Y^{(\lambda )}(\widehat{๐ซ}_2)+B.`$ (14)
There are 8 parameters in the Hamiltonian, $`E_0`$, $`ฯต_j`$ for the five orbits, $`A_{T=1}`$ and $`B`$. They can be classified into two groups. One is comprised of the differences of $`ฯต_j`$โs (4 parameters) and $`A_{T=1}`$. These five parameters are relevant to the excitation spectra for an individual nucleus. The other consists of $`E_0`$, $`B`$ and overall shift of $`ฯต_j`$โs. They do not change excitation spectra, but affect the gross behavior of the binding energies. It is noticed that effects of the Coulomb repulsion between protons are principally contained in $`B`$. As is proven in Appendix, the ExSRF (8) becomes exact for the $`10^+`$ decay with the present seniority-truncated model space and the interaction.
In describing the extended seniority isomerism, it is important to reproduce the degree of the pair excitation out of the $`Z=64`$ core. In <sup>147</sup>Tb and <sup>149</sup>Ho, a $`5/2^+`$ and a $`7/2^+`$ levels have been observed at very low energies ($`E_x1`$ MeV) . These levels could be another manifestation of the core excitation. It is hard to reproduce these levels without including the $`0g_{7/2}`$ and $`1d_{5/2}`$ orbits. Analogously, <sup>145</sup>Eu has low-lying ($`E_x1`$ MeV) states with $`11/2^{}`$, $`1/2^+`$ and $`3/2^+`$. The coupling constant $`A_{T=1}`$ and the $`ฯต_j`$ differences are determined so as to reproduce the lowest levels of $`E_x1`$ MeV in <sup>145</sup>Eu and <sup>147</sup>Tb, as well as the $`E_x3`$ MeV low-lying levels of <sup>146</sup>Gd. The adopted value of $`A_{T=1}`$ is 0.210 MeV. The results of the fitting are depicted in Fig. 1, together with several higher-lying levels, in comparison with the experimental data.
There are a few levels which are not described by the calculation. The $`3^{}`$ state of <sup>146</sup>Gd has been interpreted as an octupole collective mode including the neutron excitations . Therefore this state has been excluded from the fitting. The $`9/2^{}`$, $`7/2^{}`$ and $`13/2^{}`$ states of <sup>145</sup>Eu are considered to be $`\pi d_{5/2}^13^{}`$ or $`\pi g_{7/2}^13^{}`$ . Since they involve the octupole collective excitation, these states are beyond the model space in the present calculation as well. The $`15/2^+`$ and $`17/2^+`$ states of <sup>147</sup>Tb are also regarded as $`\pi h_{11/2}3^{}`$.
The remaining parameters, $`E_0`$, $`B`$ and the constant shift of the single-particle energies, are fixed from the binding energies of the $`63Z74`$, $`N=82`$ isotones . We obtain $`E_0=90.10`$ MeV and $`B=0.409`$ MeV, representing the energies by relative values to the experimental ground state energy of <sup>146</sup>Gd. The resultant single-particle energies are listed in Table I. The calculated binding energies are compared with the data in Fig. 2. We have sufficiently good agreement, with the largest discrepancy of 0.45 MeV for <sup>145</sup>Eu.
The ground-state wave function of <sup>146</sup>Gd holds the $`Z=64`$ closure only by 11% in this calculation. The core is broken due to the pairing correlation, with keeping the seniority a good quantum number. Having 53% excitation of a single pair and 36% of two pairs, average number of protons excited out of the $`Z=64`$ core amounts to 2.5. This result is barely influenced even if we relax the seniority truncation to $`v4`$. As was pointed out in Ref. , the $`Z=64`$ core is broken to a sizable extent by the pair excitation.
We carry out a shell model calculation with the above Hamiltonian for the $`66Z72`$, $`N=82`$ nuclei. The calculated energy levels for the even-$`Z`$ nuclei are compared with the observed ones in Figs. 3 (for even-parity levels) and 4 (for odd-parity levels), up to $`E_x3`$ MeV. Almost all levels in this energy range are in reasonably good agreement. Among them, the $`E_x(2^+)`$ values are somewhat higher than the data. This discrepancy seems mainly concerned with the quadrupole collectivity, and could be ascribed to the truncated model space or to the interaction which might be too simple. On the other hand, the $`E2`$ decay of the isomers has non-collective character, occurring via the transition within $`0h_{11/2}`$. Therefore it is not quite relevant to the quadrupole collectivity. As presented in Fig. 4, the odd-parity levels are also reproduced, except for the octupole collective state $`3^{}`$, which is not shown in the figure. This is an obvious advantage over against the previous single-$`j`$ calculation, since the odd-parity levels are out of the model space in the single-$`j`$ calculation. As mentioned in Section II, the excitation energies of $`2^+`$, $`8^+`$ and $`10^+`$ states slightly decrease as $`Z`$ goes up. This behavior is well reproduced by the present calculation, while the energies slightly increase in the single-$`j`$ model.
In Figs. 5 and 6, the calculated yrast levels are compared with the experimental data for the odd-$`Z`$ nuclei, up to $`E_x`$3 MeV. The energies relative to the $`11/2^{}`$ state are presented both for the data and the calculated results. The agreement is sufficiently good, as in the even-$`Z`$ nuclei. The even-parity states, which are beyond the space in the single-$`j`$ model with $`0h_{11/2}`$, are also reproduced (Fig. 6). In all of the calculated levels presented in the figures, the seniority is conserved to an excellent extent. The $`11/2^{}`$, $`1/2^+`$, $`3/2^+`$, $`5/2^+`$ and $`7/2^+`$ states lying in $`E_x1`$ MeV have $`v=1`$, while the others have $`v=3`$. It should be remarked that the $`5/2^+`$ and $`7/2^+`$ levels, the $`1d_{5/2}`$ and $`0g_{7/2}`$ states with the pair excitation, are also reproduced well, in <sup>149</sup>Ho and <sup>151</sup>Tm. The intruder level $`15/2^+`$ is not shown in the figure, which should be an octupole collective state with $`\pi 0h_{11/2}3^{}`$.
### C $`E2`$ strengths of the high-spin isomers
Let us turn to the $`E2`$ transition strengths of the high-spin isomers. The $`E2`$ operator is given by
$$T(E2)=e_{\mathrm{eff}}\underset{j,j^{}}{}\frac{1}{\sqrt{5}}j^{}r^2Y^{(2)}(\widehat{๐ซ})j[a_j^{}^{}\stackrel{~}{a}_j]^{(2)},$$
(15)
where $`\stackrel{~}{a}_{jm}=()^{j+m}a_{jm}`$. The single-particle matrix element $`j^{}r^2Y^{(2)}(\widehat{๐ซ})j`$ is evaluated by using the harmonic oscillator single-particle wave functions with the oscillator parameter $`\nu (=1/b^2)=M\omega /\mathrm{}=0.98A^{1/3}\mathrm{fm}^2`$.
It should be noticed that, in the $`E2`$ calculation, there remains only a single adjustable parameter $`e_{\mathrm{eff}}`$, the effective charge. It is found that $`e_{\mathrm{eff}}=2.3e`$ fits well to all of the $`10^+`$ and $`27/2^{}`$ decays. This value is significantly larger than the effective charge of $`1.5e`$ which was adopted in the single-$`j`$ calculation . This is in contrast to the collective transitions, where $`e_{\mathrm{eff}}`$ should be smaller as the model space is extended, since the matrix elements of $`T(E2)/e_{\mathrm{eff}}`$ tend to increase. For the $`E2`$ transitions of the isomers, which do not have collective character, the matrix elements of $`T(E2)/e_{\mathrm{eff}}`$ are smaller in the multi-$`j`$ case than in the single-$`j`$ case at <sup>148</sup>Dy and <sup>149</sup>Ho, as is recognized from Eq. (8).
While the effective charge of $`1.4e`$ was recommended in realistic calculations in the $`Z<64`$ region , several calculations in the $`Z64`$ region assumed $`e_{\mathrm{eff}}=(2.02.25)e`$ . The origin of the difference in the effective charge between $`Z<64`$ and $`Z64`$ is not clear, because in either case the model space consists of all the five orbits in the $`50<Z<82`$ shell, and the orbital-dependence of the effective charge is normally weak . We just point out that our value seems consistent with those of the previous studies in the $`Z64`$ region.
In Fig. 7, we show the $`B(E2;10^+8^+)`$ values for the $`66Z72,N=82`$ isotones. The calculated values are compared with the measured ones, as well as with those obtained in the single-$`j`$ calculation by Lawson . The $`E2`$ hindrance at $`Z=70`$ (i.e. <sup>152</sup>Yb) occurs also in the present multi-$`j`$ calculation. Our calculation gives $`B(E2)=0.6e^2\mathrm{fm}^4`$, in good agreement with the data $`0.9\pm 0.1e^2\mathrm{fm}^4`$ .
As has been shown by the ExSRF (8), the $`E2`$ strengths of the $`10^+`$ states are essentially determined from the occupation number $`N_{0h_{11/2}}`$. In the present calculation, the wave functions of the $`10^+`$ state and of $`8^+`$ yield $`N_{0h_{11/2}}=5.7`$ in <sup>152</sup>Yb. This occupation number close to $`\mathrm{\Omega }_{0h_{11/2}}=6`$ gives rise to the strong $`E2`$ hindrance. We view this hindrance from another standpoint. See Eq. (6), recalling $`\xi =0h_{11/2}`$. By decomposing the wave functions as in Eqs. (2) and (3), we look into the contribution of each $`n_\xi `$ component. Table II illustrates $`_\alpha c_{n_\xi \alpha }^2`$, $`\xi ^{n_\xi }v_\xi =vJ_f^\pi T(E2)\xi ^{n_\xi }v_\xi =vJ_i^\pi `$ and their product, for each $`n_\xi (=2,4,6,8,10)`$. As the SRF tells us, the matrix element $`\xi ^{n_\xi }v_\xi =vJ_f^\pi T(E2)\xi ^{n_\xi }v_\xi =vJ_i^\pi `$ changes its sign at $`n_\xi =6`$, where it vanishes. The coefficient $`_\alpha c_{n_\xi \alpha }^2`$ has the same sign and the same order of magnitude between $`n_\xi `$ and $`12n_\xi `$, causing a large cancellation. As a result, the $`E2`$ strength is significantly hindered for <sup>152</sup>Yb. Although the single-$`j`$ picture discussed by Lawson does not apply anymore, this mechanism, $`N_{0h_{11/2}}6`$ or in other words the cancellation of the matrix elements, explains why the $`E2`$ hindrance occurs in <sup>152</sup>Yb. Thus the $`10^+`$ state of <sup>152</sup>Yb yields a typical example of the extended seniority isomerism.
The $`E2`$ strengths in the other even-$`Z`$ isotones are also in remarkably good agreement with the data. We clearly view improvement over the single-$`j`$ model in <sup>154</sup>Hf.
As is viewed in Fig. 3, the $`4^+`$ and $`6^+`$ states have not yet been detected in <sup>152</sup>Yb and <sup>154</sup>Hf. The ExSRF (8) approximately applies also to the $`8^+6^+`$ and $`6^+4^+`$ $`E2`$ transitions. These transitions are hindered by the same mechanism as in the $`10^+8^+`$ transition. Hence it is not easy to populate the $`4^+`$ and $`6^+`$ states in the experiments.
The $`E2`$ strengths of the $`27/2^{}`$ states are shown in Fig. 8, for the $`67Z71,N=82`$ isotones. As in the $`10^+`$ isomers in the even-$`Z`$ isotones, the present calculation reproduces the measured values remarkably well. The hindrance at $`Z=71`$ (i.e. <sup>153</sup>Lu), which was not described well in the single-$`j`$ model, is reproduced. In the light of the ExSRF, this hindrance occurs because of $`N_{0h_{11/2}}=6.2`$ in <sup>153</sup>Lu. It should be emphasized that the $`E2`$ properties of the isomers are naturally reproduced, by adjusting the energies relevant to the excitation out of the $`Z=64`$ core.
Figure 9 depicts the occupation number $`N_{0h_{11/2}}`$ in the $`10^+`$ or $`27/2^{}`$ isomers, which corresponds to their $`E2`$ decay strengths via the ExSRF. We view almost linear increase of $`N_{0h_{11/2}}`$ in the isomers, in coincidence with schematic illustration by Blomqvist (Fig. 3-2 of Ref. ). The number of the particles excited out of the $`Z=64`$ core $`N_{\mathrm{exc}}14(N_{0g_{7/2}}+N_{1d_{5/2}})`$ in the isomers is plotted as well, in the right panel of Fig. 9. It is found that the number of the excited particles diminishes only gradually, as $`Z`$ increases.
## V Discussion โ Necessity of $`Z=64`$ core excitation
The ExSRF derived in Section III accounts for the seniority mechanism to hinder the $`E2`$ decay of a certain class of isomers. In Section IV, we have demonstrated that the strong $`E2`$ hindrance in <sup>152</sup>Yb is reproduced by taking the $`Z=64`$ core excitation into account. However, the ExSRF itself does not exclude the possibility of the single-$`j`$ solution to the hindrance for <sup>152</sup>Yb. In this section we argue that the $`E2`$ properties of the isomers exclusively indicate the presence of the excitation across $`Z=64`$.
According to the ExSRF (8), the vanishing $`E2`$ strength at <sup>152</sup>Yb indicates $`N_{0h_{11/2}}6`$. In respect to the stiffness of the $`Z=64`$ core, the following two possibilities result: (i) $`\pi 0h_{11/2}`$ couples to the surrounding orbits very weakly and the single-$`j`$ picture holds to a good approximation, or (ii) the pair excitation across $`Z=64`$ compensates the pairing mixing of $`0h_{11/2}`$ with $`(2s_{1/2}1d_{3/2})`$, the possibility first suggested by Blomqvist . We have shown in Section IV that (ii) is plausible, by reproducing the energy levels and the $`E2`$ strengths simultaneously. We here discuss whether (i) is possible or not.
For this purpose we consider the $`E2`$ strength of <sup>152</sup>Yb in the $`3j`$ model of $`0h_{11/2}`$, $`2s_{1/2}`$ and $`1d_{3/2}`$, keeping the $`Z=64`$ closure. The major point will be the amount of mixing of $`0h_{11/2}`$ with the surrounding orbits $`2s_{1/2}`$ and $`1d_{3/2}`$ due to the pairing interaction. The possibility (i) requires that the mixing should be negligibly small. The valence particle number $`n`$ is 6 for <sup>152</sup>Yb in the $`3j`$ model. In order for the strong hindrance to be reproduced, the wave function of <sup>152</sup>Yb should have $`n_\xi =6`$ ($`\xi =0h_{11/2}`$) as the main component, with a small admixture of the $`n_\xi =4`$ component. By using the effective charge of Ref. , the measured $`B(E2)`$ leads to the admixture of the $`n_\xi =4`$ component by no greater than 10%.
For the sake of simplicity, let us first consider mixing between two configurations. In reality, this mixing could be either of the $`0h_{11/2}`$-$`2s_{1/2}`$ or the $`0h_{11/2}`$-$`1d_{3/2}`$ pairing mixing. The degree of the mixing is connected to the ratio of the off-diagonal matrix elements of the pairing interaction (denoted by $`V_{\mathrm{pair}}^{\mathrm{off}}`$) to the energy difference of the relevant orbits (denoted by $`\mathrm{\Delta }E`$). The mixing probability is given by $`V_{\mathrm{pair}}^{\mathrm{off}}^2/[(2\mathrm{\Delta }E)^2+V_{\mathrm{pair}}^{\mathrm{off}}^2]`$. The above 10% mixing indicates $`V_{\mathrm{pair}}^{\mathrm{off}}/2\mathrm{\Delta }E=0.45`$. If the mixing among the three orbits is considered, this ratio should be regarded as the upper limit for each of the $`0h_{11/2}`$-$`2s_{1/2}`$ and the $`0h_{11/2}`$-$`1d_{3/2}`$ mixing. The level scheme of the odd-$`Z`$ isotones <sup>147</sup>Tb, <sup>149</sup>Ho and <sup>151</sup>Lu implies that the three orbits keep nearly degenerate within the 0.2 MeV accuracy in this region. Hence we can put $`\mathrm{\Delta }E<0.2`$ MeV. We thus find that, in order for the possibility (i) to be realized, $`V_{\mathrm{pair}}^{\mathrm{off}}<0.18`$ MeV is necessary.
Generally speaking, interactions with the shorter range yield the larger off-diagonal pairing matrix elements. We estimate $`V_{\mathrm{pair}}^{\mathrm{off}}`$ in the SDI and in the Yukawa form with the range of the one-pion exchange, as representatives of short-range and long-range interactions. After fitting their strengths to the observed energy levels in the $`3j`$ model, the SDI and the Yukawa interaction give $`V_{\mathrm{pair}}^{\mathrm{off}}1`$ and 0.5 MeV, respectively . As a consequence of the weak coupling between $`0h_{11/2}`$ and $`(2s_{1/2}1d_{3/2})`$, the long-range interaction gives very low $`0_2^+`$ states, enough low to be first excited states, in <sup>148</sup>Dy and <sup>150</sup>Er. Such $`0_2^+`$ levels have not been observed. Nevertheless, even the long-range interaction gives significantly larger $`V_{\mathrm{pair}}^{\mathrm{off}}`$ than required in (i). It is practically impossible to avoid a considerable mixing of $`(2s_{1/2}1d_{3/2})`$ due to the pairing interaction. Indeed, in both calculations with the SDI and the Yukawa interaction in the $`3j`$ model space, the three orbits well mix one another via the pairing interaction. Thereby the $`E2`$ strengths of the high-spin isomers are almost described by the formula (11), the QS formula for the degenerate single-particle orbits, with $`\mathrm{\Omega }=9`$. This is obviously inconsistent with the measurement.
We thus conclude that the possibility (i) cannot be realistic. As has been discussed in Section III, in order to reproduce the $`E2`$ hindrance for <sup>152</sup>Yb the admixture of the $`2s_{1/2}`$ and $`1d_{3/2}`$ orbits must be compensated by the $`Z=64`$ core excitation. Therefore, with the nearly degeneracy between $`0h_{11/2}`$ and $`(2s_{1/2}1d_{3/2})`$ taken into consideration, the $`E2`$ properties of the $`N=82`$ isomers are an exclusive evidence for the presence of the excitation across $`Z=64`$, not indicating stiff $`Z=64`$ core. The seniority isomerism in this region is a probe sensitive to the $`Z=64`$ core excitation due to the pairing correlations. It is difficult to handle the influence of the $`Z=64`$ core breaking on the $`E2`$ properties of the isomers by renormalization. The presence of substantial pair excitation is a clear difference of the submagic number $`Z=64`$ from the ordinary magic numbers.
It is commented here that, in contrast to the SDI adopted in Section IV, Wenes et al. applied a finite-range interaction with the Gaussian form to the nuclei in this region . As a result of the weak coupling between $`0h_{11/2}`$ and $`(2s_{1/2}1d_{3/2})`$, their calculation predicted $`0_2^+`$ states at unusually low energies in <sup>148</sup>Dy and <sup>150</sup>Er, quite similar to the above $`3j`$ case with the Yukawa interaction.
Wildenthal proposed an effective Hamiltonian for the $`N=82`$ isotones . Starting from the SDI+QQ interaction, the interaction matrix elements are fitted to the $`50<Z72`$ nuclei. While Wildenthalโs Hamiltonian (with the assumed truncation) nicely describes the energy levels, it does not reproduce the $`Z`$-dependence of the $`E2`$ properties of the isomers; in particular, the strong hindrance around <sup>152</sup>Yb. This is because the pair excitation out of the $`Z=64`$ core is too small, at least for the high-spin isomers.
The above possibilities (i) and (ii) can also be judged by future experiments. Though the ExSRF (8) connects the $`E2`$ strength to the occupation number $`N_{0h_{11/2}}`$, the ambiguity in $`e_{\mathrm{eff}}`$ prohibits us from extracting $`N_{0h_{11/2}}`$ directly from $`B(E2)`$ of the isomers, in practice. The two possibilities (i) and (ii) give somewhat similar $`E2`$ strengths for the high-spin isomers on account of the difference in $`e_{\mathrm{eff}}`$, in the discussions so far. However, this does not apply to the transition from the lowest $`2^+`$ state to the ground $`0^+`$ state. In Table III, the $`B(E2;2^+0^+)`$ values calculated in the multi-$`j`$ model are compared with those in the single-$`j`$ model. Without state-dependence of the effective charges, we predict $`2.55`$ times larger $`E2`$ strengths in the multi-$`j`$ space (i.e. (ii)) than in the single-$`j`$ space (i.e. (i)). In the multi-$`j`$ model of Section IV, there seems to be a problem with respect to the quadrupole collectivity. Hence we should not expect an excellent precision on the multi-$`j`$ prediction; indeed, by a slight variation of the interaction $`B(E2;2^+0^+)`$ can deviate by 30% without influencing the $`E2`$ strengths of the isomers. Still the big difference between the single-$`j`$ and multi-$`j`$ models would enable us to judge which of the two possibilities (i) or (ii) is reliable.
## VI Summary and outlook
We have investigated the $`10^+`$ and $`27/2^{}`$ isomers of the $`Z>64`$, $`N=82`$ nuclei. The extended seniority reduction formula has been derived for the $`E2`$ decay strengths of the isomers, under reasonable assumptions. This formula links the $`E2`$ strength to the occupation number on the $`\pi 0h_{11/2}`$ orbit, apart from the ambiguity in the effective charge. The extended formula accounts for the mechanism of the $`E2`$ hindrance, which we have called extended seniority isomerism.
By taking into account the excitations from $`(0g_{7/2}1d_{5/2})`$ to $`(0h_{11/2}2s_{1/2}1d_{3/2})`$, the binding energies, the energy levels of both parities and the $`B(E2)`$ values have simultaneously been reproduced in a multi-$`j`$ shell model calculation with the MSDI. The $`E2`$ hindrance in <sup>153</sup>Lu as well as in <sup>152</sup>Yb has been described quite well. Combined with the approximate degeneracy among the $`0h_{11/2}`$, $`2s_{1/2}`$ and $`1d_{3/2}`$ orbits, the strong $`E2`$ hindrance around $`Z=70`$ exclusively indicates the presence of the pair excitation out of the $`Z=64`$ core. Thus the $`Z=64`$ core is not very stiff. It is not always justified to assume the <sup>146</sup>Gd inert core, even for the relatively low-lying states in the $`N=82`$ isotones. In this respect, the number $`Z=64`$ should be distinguished from the magic numbers like $`N=82`$, though it could be fair to be called submagic number.
The extended seniority isomerism may exist in other single-closed nuclei and their neighbors. While we have restricted our discussion to the $`Z>64`$, $`N=82`$ nuclei, focusing on the stiffness of the $`Z=64`$ core, it is of interest to apply similar approaches to nuclei in other mass region. Work in this line is under progress.
This work was supported in part by the Ministry of Education, Science, Sports and Culture of Japan (Grant-in-Aid for Encouragement of Young Scientists, No. 11740137).
## Appendix: Argument on the condition (c) in Section III
The condition (c) in Section III is expected to be a good approximation. Indeed, it is exactly derived from (a) and (b) if the interaction within the $`\xi `$-subspace has the zero-range character. We prove it in this Appendix.
### A General argument
With assuming the conditions (a) and (b), let us consider matrix elements of the general shell model Hamiltonian between the bases appearing in Eqs. (2) and (3). We shall first prove the relation
$`\left(\xi ^{n_\xi ^{}}v_\xi =vJ_f^\pi |r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\right)H\left(|\xi ^{n_\xi }v_\xi =vJ_f^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+\right)`$ (16)
$`\left(\xi ^{n_\xi ^{}}v_\xi =vJ_i^\pi |r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\right)H\left(|\xi ^{n_\xi }v_\xi =vJ_i^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+\right)`$ (17)
$`=\delta _{n_\xi ,n_\xi ^{}}\delta _{\alpha ,\alpha ^{}}\left(\xi ^{n_\xi }vJ_f^\pi |V_\xi |\xi ^{n_\xi }vJ_f^\pi \xi ^{n_\xi }vJ_i^\pi |V_\xi |\xi ^{n_\xi }vJ_i^\pi \right),`$ (18)
as far as the Hamiltonian consists of single-particle energies and of two-body residual interaction. Here $`V_\xi `$ stands for the two-body interaction within the $`\xi `$ subspace. The left-hand side (LHS) of Eq. (18) obviously vanishes for the single-particle energy term of $`H`$. It is sufficient to focus on matrix elements of the two-body interaction.
The two-body interaction is expressed, in the second-quantized form, by the sum of the terms composed of $`a_{j_1}^{}a_{j_2}^{}a_{j_3}a_{j_4}`$ operators (with coupling constants). According to which of the $`j`$โs belong to $`\xi `$ or $`r`$, all the possible terms contributing to the matrix elements are classified into the following categories: (i) $`a_\xi ^{}a_\xi ^{}a_ra_r`$ terms and their hermitian conjugates, (ii) $`a_r^{}a_r^{}a_ra_r`$ terms, (iii) $`a_\xi ^{}a_r^{}a_ra_\xi `$ terms, and (iv) $`a_\xi ^{}a_\xi ^{}a_\xi a_\xi `$ terms. It is noted that, since $`v_r=0`$, the terms having an odd number of $`(a_r^{},a_r)`$ operators vanish. We decompose the operators into the $`\xi `$ part and the $`r`$ part, and denote them by $`\widehat{h}_\xi `$ and $`\widehat{h}_r`$. The matrix elements are also decomposed into the $`\xi `$ and $`r`$ part,
$`\left(\xi ^{n_\xi ^{}}v_\xi =vJ^\pi |r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\right)\widehat{h}_\xi \widehat{h}_r\left(|\xi ^{n_\xi }v_\xi =vJ^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+\right)`$ (19)
$`=\xi ^{n_\xi ^{}}v_\xi =vJ^\pi |\widehat{h}_\xi |\xi ^{n_\xi }v_\xi =vJ^\pi r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\widehat{h}_r|r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+.`$ (20)
It is obvious that the $`r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\widehat{h}_r|r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+`$ part does not depend on $`J^\pi `$. Therefore,
$`\left(\xi ^{n_\xi ^{}}v_\xi =vJ_f^\pi |r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\right)\widehat{h}_\xi \widehat{h}_r\left(|\xi ^{n_\xi }v_\xi =vJ_f^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+\right)`$ (21)
$`\left(\xi ^{n_\xi ^{}}v_\xi =vJ_i^\pi |r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\right)\widehat{h}_\xi \widehat{h}_r\left(|\xi ^{n_\xi }v_\xi =vJ_i^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+\right)`$ (22)
$`=r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\widehat{h}_r|r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+`$ (23)
$`\times \left(\xi ^{n_\xi ^{}}v_\xi =vJ_f^\pi |\widehat{h}_\xi |\xi ^{n_\xi }v_\xi =vJ_f^\pi \xi ^{n_\xi ^{}}v_\xi =vJ_i^\pi |\widehat{h}_\xi |\xi ^{n_\xi }v_\xi =vJ_i^\pi \right).`$ (24)
For the matrix elements between the $`v_r=0`$ bases, $`\widehat{h}_r`$ cannot carry angular momentum, and therefore $`\widehat{h}_\xi `$ cannot either. Then $`\widehat{h}_\xi `$ and $`\widehat{h}_r`$ for each category are defined as, without loss of generality, (i) $`\widehat{h}_\xi =[a_\xi ^{}a_\xi ^{}]^{(0)}`$, $`\widehat{h}_r=[\stackrel{~}{a}_r\stackrel{~}{a}_r]^{(0)}`$, (ii) $`\widehat{h}_\xi =1`$, $`\widehat{h}_r=[a_r^{}a_r^{}\stackrel{~}{a}_r\stackrel{~}{a}_r]^{(0)}`$, (iii) $`\widehat{h}_\xi =[a_\xi ^{}\stackrel{~}{a}_\xi ]^{(0)}`$, $`\widehat{h}_r=[a_r^{}\stackrel{~}{a}_r]^{(0)}`$, and (iv) $`\widehat{h}_\xi =[a_\xi ^{}a_\xi ^{}\stackrel{~}{a}_\xi \stackrel{~}{a}_\xi ]^{(0)}`$, $`\widehat{h}_r=1`$. $`V_\xi `$ in Eq. (18) represents the collection of the $`\widehat{h}_\xi `$โs belonging to (iv). We discuss the matrix elements of $`\widehat{h}_\xi `$ in the right-hand side (RHS) of Eq. (24), respective to the above four categories.
The category (i) leads to $`n_\xi ^{}=n_\xi \pm 2`$ off-diagonal elements. The $`\widehat{h}_\xi =[a_\xi ^{}a_\xi ^{}]^{(0)}`$ operator in this case is proportional to a generator of the quasi-spin in the orbit $`\xi `$. Thus $`\xi ^{n_\xi +2}v_\xi =vJ^\pi |\widehat{h}_\xi |\xi ^{n_\xi }v_\xi =vJ^\pi `$ depends only on $`n_\xi `$ and $`v_\xi `$, not on $`J^\pi `$. Namely,
$$\xi ^{n_\xi ^{}}v_\xi =vJ_f^\pi |[a_\xi ^{}a_\xi ^{}]^{(0)}|\xi ^{n_\xi }v_\xi =vJ_f^\pi =\xi ^{n_\xi ^{}}v_\xi =vJ_i^\pi |[a_\xi ^{}a_\xi ^{}]^{(0)}|\xi ^{n_\xi }v_\xi =vJ_i^\pi ,$$
(25)
and the RHS of Eq. (24) vanishes.
Since $`\widehat{h}_\xi =1`$ in the category (ii), the relevant matrix elements of the $`\xi `$ part are
$$\xi ^{n_\xi ^{}}v_\xi =vJ_f^\pi |\xi ^{n_\xi }v_\xi =vJ_f^\pi =\xi ^{n_\xi ^{}}v_\xi =vJ_i^\pi |\xi ^{n_\xi }v_\xi =vJ_i^\pi =\delta _{n_\xi ,n_\xi ^{}}.$$
(26)
For the category (iii), $`\widehat{h}_\xi =[a_\xi ^{}\stackrel{~}{a}_\xi ]^{(0)}N_\xi `$, leading to
$$\xi ^{n_\xi ^{}}v_\xi =vJ_f^\pi |[a_\xi ^{}\stackrel{~}{a}_\xi ]^{(0)}|\xi ^{n_\xi }v_\xi =vJ_f^\pi =\xi ^{n_\xi ^{}}v_\xi =vJ_i^\pi |[a_\xi ^{}\stackrel{~}{a}_\xi ]^{(0)}|\xi ^{n_\xi }v_\xi =vJ_i^\pi \delta _{n_\xi ,n_\xi ^{}}n_\xi .$$
(27)
Both terms do not contribute to the RHS of Eq. (24).
From Eqs. (25), (26) and (27), only the terms of the category (iv) may contribute to the RHS of Eq. (24). Equation (18) follows, because $`\widehat{h}_r=1`$ for the category (iv), with replacing the sum of $`\widehat{h}_\xi `$โs by $`V_\xi `$. The argument is now reduced to the single-$`j`$ matrix elements within the $`\xi `$ orbit. Remark again that we have not imposed any restriction on the Hamiltonian in the discussion so far, besides that it consists of the single-particle energies and two-body interaction.
### B Property of $`V_\xi `$
We next consider the property of $`V_\xi `$. If $`\xi ^{n_\xi }vJ^\pi |V_\xi |\xi ^{n_\xi }vJ^\pi `$ is independent of $`n_\xi `$ or $`J`$, we have
$`\xi ^{n_\xi }vJ_f^\pi |V_\xi |\xi ^{n_\xi }vJ_f^\pi \xi ^{n_\xi }vJ_i^\pi |V_\xi |\xi ^{n_\xi }vJ_i^\pi `$ (28)
$`=\xi ^vvJ_f^\pi |V_\xi |\xi ^vvJ_f^\pi \xi ^vvJ_i^\pi |V_\xi |\xi ^vvJ_i^\pi .`$ (29)
Substituting it into the RHS of Eq. (18), we obtain
$`\left(\xi ^{n_\xi ^{}}v_\xi =vJ_f^\pi |r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\right)H\left(|\xi ^{n_\xi }v_\xi =vJ_f^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+\right)`$ (30)
$`\left(\xi ^{n_\xi ^{}}v_\xi =vJ_i^\pi |r^{nn_\xi ^{}}\alpha ^{}v_r=\mathrm{0\; 0}^+|\right)H\left(|\xi ^{n_\xi }v_\xi =vJ_i^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+\right)`$ (31)
$`=\delta _{n_\xi ,n_\xi ^{}}\delta _{\alpha ,\alpha ^{}}\left(\xi ^vvJ_f^\pi |V_\xi |\xi ^vvJ_f^\pi \xi ^vvJ_i^\pi |V_\xi |\xi ^vvJ_i^\pi \right).`$ (32)
The Hamiltonian matrix can be separated according to $`J`$, because of the angular momentum conservation. Moreover, since the seniority $`v`$ has been assumed to be a good quantum number, it is sufficient to consider submatrices of $`H`$ for a fixed $`v`$. The space to be diagonalized is spanned by $`|\xi ^{n_\xi }v_\xi =vJ^\pi |r^{nn_\xi }\alpha v_r=\mathrm{0\; 0}^+`$ with various $`n_\xi `$ and $`\alpha `$ (see Eqs. (2) and (3)). Equation (32) implies that the submatrices of $`H`$ are identical between $`|(\xi r)^nvJ_i^\pi `$ and $`|(\xi r)^nvJ_f^\pi `$, except for a constant shift of the diagonal elements. Diagonalized by the same unitary matrix, the lowest eigenstates $`|J_i^\pi `$ and $`|J_f^\pi `$ have equal coefficient $`c_{n_\xi \alpha }`$ to each other; the condition (c) is exactly satisfied. It is now clear that Eq. (29) is crucial to the condition (c).
In the quasi-spin (QS) regime within the single orbit $`\xi `$, $`[a_\xi ^{}a_\xi ^{}\stackrel{~}{a}_\xi \stackrel{~}{a}_\xi ]^{(0)}`$ can be QS-scalar, vector or tensor, in general. If $`V_\xi `$ is purely QS-scalar, the matrix element $`\xi ^{n_\xi }v_\xi =vJ_i^\pi |V_\xi |\xi ^{n_\xi }v_\xi =vJ_i^\pi `$ is independent of $`n_\xi `$, and Eq. (29) is fulfilled. Equation (29) is also satisfied if QS-vector and QS-tensor parts of $`V_\xi `$ are $`J`$-independent. This is indeed attained when the QS-vector and tensor parts can be expressed by the QS generators ($`[a_\xi ^{}a_\xi ^{}]^{(0)}`$, $`[\stackrel{~}{a}_\xi \stackrel{~}{a}_\xi ]^{(0)}`$ and $`[a_\xi ^{}\stackrel{~}{a}_\xi ]^{(0)}`$, besides appropriate constant factors). An immediate example is the monopole pairing ($`[a_\xi ^{}a_\xi ^{}]^{(0)}[\stackrel{~}{a}_\xi \stackrel{~}{a}_\xi ]^{(0)}`$). A sufficient condition to Eq. (29) is that $`V_\xi `$ consists only of QS-scalars and of the QS generators.
The general form of $`V_\xi `$ can be represented by
$$V_\xi =\underset{\lambda =\mathrm{even}}{}\frac{g_\lambda }{2}[a_\xi ^{}a_\xi ^{}]^{(\lambda )}[\stackrel{~}{a}_\xi \stackrel{~}{a}_\xi ]^{(\lambda )}.$$
(33)
The corresponding โparticle-holeโ interaction is defined by
$$\overline{V}_\xi =\underset{\lambda }{}f_\lambda [a_\xi ^{}a_\xi ^{}]^{(\lambda )}[\stackrel{~}{a}_\xi \stackrel{~}{a}_\xi ]^{(\lambda )},$$
(34)
with
$$f_\lambda =\underset{\lambda ^{}=\mathrm{even}}{}(2\lambda ^{}+1)W(j_\xi j_\xi j_\xi j_\xi ;\lambda ^{}\lambda )g_\lambda ^{}.$$
(35)
In Eq. (34), only $`\lambda =\mathrm{even}`$ terms remain owing to the antisymmetrization. According to the QS argument , we have
$`\xi ^{n_\xi }vJ^\pi |V_\xi |\xi ^{n_\xi }vJ^\pi `$ $`=`$ $`\left\{{\displaystyle \frac{(\mathrm{\Omega }_\xi 2v)(2\mathrm{\Omega }_\xi n_\xi v)}{4(\mathrm{\Omega }_\xi v)(\mathrm{\Omega }_\xi v1)}}(g_0+2f_0)f_0\right\}(n_\xi v)`$ (38)
$`+{\displaystyle \frac{(\mathrm{\Omega }_\xi v)(\mathrm{\Omega }_\xi v2)+(n_\xi \mathrm{\Omega }_\xi )^2}{2(\mathrm{\Omega }_\xi v)(\mathrm{\Omega }_\xi v1)}}\xi ^vvJ^\pi |V_\xi |\xi ^vvJ^\pi `$
$`+{\displaystyle \frac{(\mathrm{\Omega }_\xi v)^2(n_\xi \mathrm{\Omega }_\xi )^2}{2(\mathrm{\Omega }_\xi v)(\mathrm{\Omega }_\xi v1)}}\xi ^vvJ^\pi |\overline{V}_\xi |\xi ^vvJ^\pi .`$
By subtracting out the $`n_\xi `$\- and $`J`$-independent terms, Eq. (38) derives
$`\left(\xi ^{n_\xi }vJ_f^\pi |V_\xi |\xi ^{n_\xi }vJ_f^\pi \xi ^{n_\xi }vJ_i^\pi |V_\xi |\xi ^{n_\xi }vJ_i^\pi \right)`$ (39)
$`\left(\xi ^vvJ_f^\pi |V_\xi |\xi ^vvJ_f^\pi \xi ^vvJ_i^\pi |V_\xi |\xi ^vvJ_i^\pi \right)`$ (40)
$`={\displaystyle \frac{(n_\xi \mathrm{\Omega }_\xi )^2}{2(\mathrm{\Omega }_\xi v)(\mathrm{\Omega }_\xi v1)}}\left\{\xi ^vvJ_f^\pi |(V_\xi \overline{V}_\xi )|\xi ^vvJ_f^\pi \xi ^vvJ_i^\pi |(V_\xi \overline{V}_\xi )|\xi ^vvJ_i^\pi \right\}.`$ (41)
The $`n_\xi `$-dependence is eliminated if we have
$$\xi ^vvJ_f^\pi |(V_\xi \overline{V}_\xi )|\xi ^vvJ_f^\pi =\xi ^vvJ_i^\pi |(V_\xi \overline{V}_\xi )|\xi ^vvJ_i^\pi .$$
(42)
Equations (33) and (34) lead to
$$V_\xi \overline{V}_\xi =\underset{\lambda =\mathrm{even}}{}\frac{g_\lambda +2f_\lambda }{2}[a_\xi ^{}a_\xi ^{}]^{(\lambda )}[\stackrel{~}{a}_\xi \stackrel{~}{a}_\xi ]^{(\lambda )}.$$
(43)
As far as the seniority is not large, only a limited number of $`\lambda `$โs in Eq. (43) contribute to the RHS of Eq. (41). For instance, only the $`\lambda =J_i`$ and $`J_f`$ terms are relevant to the $`v=2`$ case, and Eq. (42) then derives $`g_{J_i}+2f_{J_i}=g_{J_f}+2f_{J_f}`$. When $`V_\xi `$ is QS-scalar, $`V_\xi =\overline{V}_\xi `$ and $`g_\lambda +2f_\lambda =0`$ for any even $`\lambda `$.
### C Zero-range interaction
We here verify that Eq. (29) is exactly fulfilled if $`V_\xi `$ is a zero-range interaction.
The one-body operator $`[a_\xi ^{}\stackrel{~}{a}_\xi ]^{(\lambda )}`$ is QS-scalar for an odd $`\lambda `$ . An easy way to construct a QS-scalar interaction is to take
$$\underset{\lambda =\mathrm{odd}}{}q_\lambda [a_\xi ^{}\stackrel{~}{a}_\xi ]^{(\lambda )}[a_\xi ^{}\stackrel{~}{a}_\xi ]^{(\lambda )},$$
(44)
where $`q_\lambda `$ is an arbitrary constant.
Suppose that $`V_\xi `$ is a zero-range interaction, which we here define as
$$V_\xi ^S=u(r_1,r_2)\delta (\widehat{๐ซ}_1\widehat{๐ซ}_2),$$
(45)
with the exchange symmetry $`u(r_1,r_2)=u(r_2,r_1)`$. The SDI adopted in the text is of this type ($`u(r_1,r_2)\delta (r_1r_2)\delta (r_1R)/R^2`$). Since $`V_\xi `$ is under discussion, the radial part of the interaction is unimportant, giving only an overall factor to the matrix elements. Expanding the angular part of $`V_\xi ^S`$ by the Legendre polynomials, we obtain
$`V_\xi ^S`$ $`=`$ $`u(r_1,r_2){\displaystyle \underset{\lambda }{}}{\displaystyle \frac{2\lambda +1}{2}}P_\lambda (\mathrm{cos}\theta _{12})`$ (46)
$`=`$ $`2\pi u(r_1,r_2){\displaystyle \underset{\lambda }{}}Y^{(\lambda )}(\widehat{๐ซ}_1)Y^{(\lambda )}(\widehat{๐ซ}_2).`$ (47)
On the other hand, the zero-range interaction given in Eq. (45) acts on the spatially symmetric two-body states. Therefore, if the two-body states are antisymmetrized, the zero-range interaction automatically picks up the spin-singlet two-body states for identical fermion systems. This leads to
$$V_\xi ^S=\frac{4}{3}(๐ฌ_1๐ฌ_2)V_\xi ^S=\frac{8\pi }{3}u(r_1,r_2)(๐ฌ_1๐ฌ_2)\underset{\lambda }{}Y^{(\lambda )}(\widehat{๐ซ}_1)Y^{(\lambda )}(\widehat{๐ซ}_2).$$
(48)
With the angular momentum recoupling, we rewrite it as
$$V_\xi ^S=\frac{8\pi }{3}u(r_1,r_2)\underset{\lambda ,\kappa }{}[Y^{(\lambda )}(\widehat{๐ซ}_1)๐ฌ_1]^{(\kappa )}[Y^{(\lambda )}(\widehat{๐ซ}_2)๐ฌ_2]^{(\kappa )}.$$
(49)
We now switch to the second-quantized representation. The equivalent one-body operator to $`[Y^{(\lambda )}(\widehat{๐ซ})๐ฌ]^{(\kappa )}`$ in the $`\xi `$ subspace is
$$\frac{1}{2\kappa +1}\xi [Y^{(\lambda )}(\widehat{๐ซ})๐ฌ]^{(\kappa )}\xi [a_\xi ^{}\stackrel{~}{a}_\xi ]^{(\kappa )}.$$
(50)
The single-particle matrix element in Eq. (50) is evaluated by
$$\xi [Y^{(\lambda )}(\widehat{๐ซ})๐ฌ]^{(\kappa )}\xi =\sqrt{2\kappa +1}(2j_\xi +1)\left\{\begin{array}{ccc}l_\xi & 1/2& j_\xi \\ \lambda & 1& \kappa \\ l_\xi & 1/2& j_\xi \end{array}\right\}l_\xi Y^{(\lambda )}(\widehat{๐ซ})l_\xi 1/2๐ฌ1/2.$$
(51)
In order for $`l_\xi Y^{(\lambda )}(\widehat{๐ซ})l_\xi `$ in the RHS not to vanish, $`\lambda `$ must be even (parity selection rule). On the other hand, owing to the symmetry of the $`9j`$-symbol in Eq. (51), the above matrix element vanishes if $`\lambda +1+\kappa `$ is odd. This is a consequence of the time reversality. Therefore, the single-particle matrix element of Eq. (51) vanishes for even $`\kappa `$. Back to Eq. (50), we find that $`[Y^{(\lambda )}(\widehat{๐ซ})๐ฌ]^{(\kappa )}`$ is QS-scalar because $`\kappa `$ is always odd. Hence $`V_\xi ^S`$ is also QS-scalar via Eq. (49). Thus, the zero-range $`V_\xi `$ has $`g_\lambda +2f_\lambda =0`$ in Eq. (43) for any possible $`\lambda `$, and therefore satisfies Eq. (29).
In reality, $`V_\xi `$ will not fully be zero-range. However, as far as the short-range interaction dominates, the matrix elements of $`V_\xi `$ are not very different from the zero-range interaction, having $`g_\lambda +2f_\lambda 0`$. The condition (c) is therefore expected to be a good approximation.
|
warning/0006/cond-mat0006109.html
|
ar5iv
|
text
|
# Ideal gas in nonextensive optimal Lagrange multipliers formalism
## I Introduction
The purpose of this work is to discuss the classical ideal gas problem in Tsallis thermostatistics within the framework of the method of optimal Lagrange multipliers (OLM) recently proposed in Ref. .
The ideal gas problem in the normalized Tsallis thermostatistics has recently been exhaustively discussed in Ref. . Correlations induced by Tsallisโ nonextensivity were analyzed in Ref. . An interesting relation, analogous to that yielding the mean energy of the (ordinary) ideal gas was thereby found. However, the ensuing results still depend upon Tsallisโ nonextensivity index $`q`$, even if one deals with a purely classical system. Here, we show that more insights can be gained into the problem if i) one takes advantage of a degree of freedom implicit in Tsallisโ formalism, i.e., choice of the entropy constant $`k_T`$, playing the role of Boltzmannโs constant $`k_B`$, and ii) uses the OLM technique in Ref. .
Tsallisโ thermostatistics \[4-18\] is by now known to offer a nonextensive generalization of traditional Boltzmann-Gibbs statistical mechanics. A key ingredient in this formalism is the introduction of a particular definition of expectation value termed the normalized $`q`$-expectation value . Actually, during the last ten years before the work in Ref. , several proposals have been made regarding definition of expectation value . No matter what definition one chooses, ordinary Boltzmann-Gibbs results are always reproduced in the extensive limit $`q1`$.
Tsallis thermostatistics involves extremization of Tsallisโ entropy
$$\frac{S_q}{k_T}=\frac{1๐๐ฑf^q(๐ฑ)}{q1},$$
(1)
by recourse to Lagrangeโs constrained variational technique. Here, $`๐ฑ`$ is a phase space element ($`N`$ particles in a $`D`$-dimensional space), $`q\mathrm{}`$ is Tsallisโ nonextensivity index, $`f`$ stands for any normalized probability density, and $`k_T`$ is the entropy constant that is akin to the celebrated Boltzmann constant appearing in traditional statistical mechanics.
The formalism of Ref. gives rise to a non-diagonal form of the Hessian associated with examining the extremum structure of the entropy and free energy through the Legendre transform procedure. In Ref. , a new approach has been advanced, in which the Hessian is diagonal. This approach was shown to enormously facilitate ascertaining what kind of extrema the Lagrange technique yields.
A quite interesting fact is to be emphasized concerning the entropy constant $`k_T`$, which is usually identified simply with Boltzmannโs $`k_B`$ in the literature: the only certified fact one can be sure of is โ$`k_Tk_B`$ for $`q1`$, which entails that there is room to choose $`k_T=k(q)`$ in a suitable way. It is seen that if one chooses (with $`\overline{Z}_q`$ the generalized partition function)
$$k_T=k_B\overline{Z}_q^{q1},$$
(2)
in conjunction with the OLM formalism , the classical harmonic oscillator determines a specific heat $`C_q=k_B`$, so that the ordinary Boltzmann-Gibbs result arises without invoking the limit $`q1`$. In addition, the OLM treatment is able to reproduce the zero-th law of thermodynamics, a goal that had previously eluded Tsallis-thermostatistics practitioners . One is then tempted to conjecture that many other ordinary results can be reproduced by Tsallis thermostatistics without invoking the limit $`q1`$. In this paper, we revisit the ideal gas problem in Tsallis thermostatistics with such a goal in mind. More precisely, we perform an OLM treatment of the problem following the canonical ensemble structure. We will show that the ordinary expression for the mean value of the energy is reproduced by Tsallis thermostatistics without the need of going to the limit $`q1`$, in contrast to the previous treatment of this problem.
It is worth noting here that, as shown in Ref. , the usual thermodynamic limit $`N\mathrm{}`$ implies the limit $`q1`$. There, the Tsallis probability distribution is deduced in a way that mimics Gibbsโ celebrated derivation of the canonical distribution for a system in contact with a heat bath . It was shown that Tsallisโ distribution naturally arises for finite heat baths, the nonextensivity index $`q`$ being related to the particle number $`N`$ that characterizes the bath. Gibbsโ canonical distribution, instead, results for infinite heat baths . This leads one to conclude that $`q`$ goes over to unity in the limit $`N\mathrm{}`$. This result was found employing the unnormalized expectation values of Curado and Tsallis (see Ref. ). Analogously, if one uses the normalized expectation values of Ref. , a similar argument leads to
$$q=\frac{DN4}{DN2}$$
(3)
for the ideal gas in $`D`$ dimensions. Since the limit $`N\mathrm{}`$ corresponds to $`q1`$, the same ordinary result is obtained from the nonextensive formalisms with various definitions of generalized expectation value (both normalized and unnormalized) in such a limit.
## II Brief review OLM formalism
A general classical treatment requires consideration of the probability density $`f(๐ฑ)`$ that maximizes Tsallisโ entropy, subject to the foreknowledge of the generalized expectation values of certain physical quantities.
Tsallisโ probability distribution is obtained by following the well known MaxEnt principle . The Tsallis-Mendes-Plastino variational treatment involves a set of Lagrange multipliers $`\lambda _j`$. The OLM technique developed in Ref. pursues an alternative path that involves a different set of Lagrange multipliers, say, $`\lambda _j^{}`$: one maximizes Tsallis entropy $`S_q`$ in Eq. (1) , subject to the modified constraints (โcenteredโ generalized expectation values) :
$`{\displaystyle ๐๐ฑf(๐ฑ)}1`$ $`=`$ $`0,`$ (4)
$`{\displaystyle ๐๐ฑf(๐ฑ)^q\left(O_j(๐ฑ)O_j_q\right)}`$ $`=`$ $`0,`$ (5)
where $`O_j(๐ฑ)`$ ($`j=1,2,\mathrm{},M`$) denote the $`M`$ relevant dynamical quantity (the observation level ). In the above, $`O_j_q`$ are defined by
$$O_j_q=\frac{๐๐ฑf^q(๐ฑ)O_j(๐ฑ)}{๐๐ฑf^q(๐ฑ)},$$
(6)
which are assumed to be a priori known. (The procedure given in Ref. employs non-centered expectation values.) The resulting probability distribution reads
$$f(๐ฑ)=\overline{Z}_q^1\left[1(1q)\underset{j}{\overset{M}{}}\lambda _j^{}\left(O_j(๐ฑ)O_j_q\right)\right]^{\frac{1}{1q}},$$
(7)
where $`\overline{Z}_q`$ stands for the generalized partition function
$$\overline{Z}_q=๐๐ฑ\left[1(1q)\underset{j}{\overset{M}{}}\lambda _j^{}\left(O_j(๐ฑ)O_j_q\right)\right]^{\frac{1}{1q}}.$$
(8)
Although the procedure originally devised in Ref. overcomes some problems posed by the old, unnormalized way of evaluating Tsallisโ generalized expectation values , it yields probability distributions that are self-referential. The resulting distribution includes the integral of the $`q`$th power of the distribution itself. This fact entails difficulties in numerical model calculations, for example. The complementary OLM treatment of Ref. surmounts such difficulties. The above-mentioned self-reference problem does not arise in Eq. (7).
It is shown in Ref. that the Lagrange multipliers $`\lambda _j`$ of the Tsallis-Mendes-Plastino procedure and the corresponding $`\lambda _j^{}`$ in OLM are connected to each other as follows:
$$\lambda _j^{}=\frac{\lambda _j}{๐๐ฑf^q(๐ฑ)}.$$
(9)
However, the genuine Lagrange multipliers are $`\lambda _j^{}`$ in OLM.
The probability density appearing in Eq. (9) is the one that maximizes the entropy $`S_q`$ which can be expressed in the alternative form
$$S_q=k_T\frac{\overline{Z}_q^{q1}1}{q1}๐๐ฑf^q(๐ฑ).$$
(10)
Also, the identical relation
$$๐๐ฑf^q(๐ฑ)=\overline{Z}_q^{1q}$$
(11)
holds, from which it follows that
$$S_q=k_T\mathrm{ln}_q\overline{Z}_q,$$
(12)
where $`\mathrm{ln}_qx=(1x^{1q})/(q1).`$
Eq. (11) allows us to rewrite Eq. (9) as
$$\lambda _j^{}=\frac{\lambda _j}{\overline{Z}_q^{1q}}.$$
(13)
However, we again stress that the basic variables in OLM are $`\lambda _j^{}`$.
Now, following Ref. , we define
$$\mathrm{ln}_qZ_q^{}=\mathrm{ln}_q\overline{Z}_q\underset{j}{}\lambda _j^{}O_j_q.$$
(14)
Introducing
$$k^{}=k_T\overline{Z}_q^{1q}$$
(15)
as in Ref. , we straightforwardly obtain
$`{\displaystyle \frac{S_q}{O_j_q}}`$ $`=`$ $`k^{}\lambda _j^{},`$ (16)
$`{\displaystyle \frac{}{\lambda _j^{}}}\left(\mathrm{ln}_qZ_q^{}\right)`$ $`=`$ $`O_j_q.`$ (17)
Eqs. (16) and (17) constitute the basic information-theoretic relations, on which statistical mechanics can be built ร la Jaynes . Here, one should remind the well known fact that in reconstructing statistical mechanics based on information theory, Jaynes could remove the concept of ensemble . In this sense, it is appropriate to emphasize that both the OLM and Tsallis-Mendes-Plastino formalisms employ identical a priori information, so that they are physically equivalent.
Notice that $`k^{}`$, as defined by Eq. (15), obeys the condition $`k^{}k_B`$ as $`q1`$. It is a condition that this constant must necessarily fulfill. (See Ref. ). Notice that if one makes use of the possibility of choosing $`k_T`$ as in (2), one obtains, on account of (15),
$$k^{}=k_B,$$
(18)
i.e., the classical results are obtained without going to the limit $`q1`$, as stated in the introduction.
Moreover, another interesting result obtained from OLM is
$$k^{}\lambda _j^{}=k_T\lambda _j,$$
(19)
which entails that the intensive variables are the same in both the OLM and Tsallis-Mendes-Plastino formalisms.
As a special instance of Eqs. (16), (17) and (19), let us discuss the canonical ensemble. In this case, only a single constraint regarding the system Hamiltonian is considered. Writing the associated Lagrange multiplier as $`\beta ^{}`$ and the generalized internal energy as $`U_q`$,
$`{\displaystyle \frac{S}{U_q}}=k^{}\beta ^{}`$ $`=`$ $`k_T\beta ={\displaystyle \frac{1}{T}},`$ (20)
$`{\displaystyle \frac{}{\beta ^{}}}\left(\mathrm{ln}_qZ_q^{}\right)`$ $`=`$ $`U_q,`$ (21)
where
$$\mathrm{ln}_qZ_q^{}=\mathrm{ln}_q\overline{Z}_q\beta ^{}U_q.$$
(22)
The temperature $`T`$ in Eq. (20) is the same as that in the Tsallis-Mendes-Plastino formalism.
## III Ideal gas in Tsallis-Mendes-Plastino formalism
The classical ideal gas in $`D`$-dimensional space in the Tsallis-Mendes-Plastino formalism has been considered in Refs. . For comparison, we recapitulate here some of its main results in the case $`0<q<1`$. The Hamiltonian reads
$$H(๐)=\underset{i=1}{\overset{N}{}}\frac{๐ฉ_i^2}{2m},$$
(23)
where $`m`$ is the particle mass, $`N`$ the particle number and $`๐ฉ_i`$ the momentum of the $`i`$th particle. We are writing the $`N`$-particle momenta collectively as $`๐=(๐ฉ_1,๐ฉ_2,\mathrm{},๐ฉ_N)`$. One extremizes the entropy in Eq. (1), subject to the constraints
$`{\displaystyle ๐\mathrm{\Omega }f(๐)}`$ $`=`$ $`1,`$ (24)
$`{\displaystyle \frac{1}{c}}{\displaystyle ๐\mathrm{\Omega }f(๐)^qH(๐)}`$ $`=`$ $`U_q,`$ (25)
where $`d\mathrm{\Omega }=(1/(N!h^{DN}))_{i=1}^Nd๐ช_id๐ฉ_i`$, with $`h`$ the linear dimension (i.e. the size) of the elementary cell in phase space, and $`_{i=1}^Nd๐ช_i=V^N`$ with the spatial volume $`V`$. We have also introduced the quantity $`c`$
$$c=๐\mathrm{\Omega }f(๐)^q.$$
(26)
The equilibrium probability density is found to be
$$f(๐)=\frac{1}{\stackrel{~}{Z}_q}\left[1(1q)(\beta /c)(H(๐)U_q)\right]^{\frac{1}{1q}},$$
(27)
where the generalized partition function is
$$\stackrel{~}{Z}_q=๐\mathrm{\Omega }\left[1(1q)(\beta /c)(H(๐)U_q)\right]^{\frac{1}{1q}},$$
(28)
and we have
$$c=[\stackrel{~}{Z}_q]^{1q}.$$
(29)
In the above, $`\beta `$ is the Lagrange multiplier associated with the constraint in Eq. (25). Notice that Eq. (27) is indeed a self-referential expression through $`c`$. In the present circumstances, however, this problem can be overcome, since a straightforward mathematical manipulation yields
$$\frac{\beta U_q}{c}=\frac{DN}{2},$$
(30)
with $`c`$ given by
$`c`$ $`=`$ $`\{{\displaystyle \frac{\mathrm{\Gamma }\left(\frac{1}{1q}\right)}{\mathrm{\Gamma }\left(\frac{1}{1q}+\frac{DN}{2}\right)}}{\displaystyle \frac{V^N}{N!h^{DN}}}\left[{\displaystyle \frac{2\pi m}{(1q)\beta }}\right]^{\frac{DN}{2}}`$ (32)
$`[1+(1q){\displaystyle \frac{DN}{2}}]^{q(1q)+DN/2}\}^{\frac{2(1q)}{2(1q)DN}}.`$
To discuss the statistical properties of the particle energies, the $`i`$th and $`j`$th single-particle Hamiltonians may be considered, namely, $`H_i=๐ฉ_i^2/2m`$ and $`H_j=๐ฉ_j^2/2m`$, respectively. Their generalized variance, covariance and correlation coefficient are defined by
$`(\mathrm{\Delta }_qH_i)^2`$ $`=`$ $`H_i{}_{}{}^{2}_qH_i_q^2,`$ (33)
$`C_q(H_i,H_j)`$ $`=`$ $`H_iH_j_qH_i_qH_j_q,`$ (34)
$`\rho (H_i,H_j)`$ $`=`$ $`{\displaystyle \frac{C_q(H_i,H_j)}{\sqrt{(\mathrm{\Delta }_qH_i)^2(\mathrm{\Delta }_qH_j)^2}}},`$ (35)
respectively. A straightforward calculation shows that these quantities are given by
$`(\mathrm{\Delta }_qH_i)^2`$ $`=`$ $`{\displaystyle \frac{c^2}{2\beta ^2}}{\displaystyle \frac{2D+(1q)D^2(N1)}{42q(1q)DN}},`$ (36)
$`C_q(H_i,H_j)`$ $`=`$ $`{\displaystyle \frac{c^2}{2\beta ^2}}{\displaystyle \frac{(1q)D^2}{42q+(1q)DN}},`$ (37)
$`\rho (H_i,H_j)`$ $`=`$ $`{\displaystyle \frac{(1q)D}{2+(1q)D(N1)}}.`$ (38)
In the limit $`q1`$, one gets
$`(\mathrm{\Delta }_qH_i)^2`$ $``$ $`{\displaystyle \frac{D}{2}}\left({\displaystyle \frac{h^2}{2\pi m}}\right)^2e^{(2+4/D)}\left({\displaystyle \frac{N}{V}}\right)^{4/D},`$ (39)
$`C_q(H_i,H_j)`$ $``$ $`0,`$ (40)
$`\rho (H_i,H_j)`$ $``$ $`0.`$ (41)
## IV Ideal gas in OLM formalism
We revisit here the classical ideal gas problem considered in the previous section within the OLM framework. We extremize Eq. (1), subject now to the constraints
$`{\displaystyle ๐\mathrm{\Omega }f(๐)}1`$ $`=`$ $`0`$ (42)
$`{\displaystyle ๐\mathrm{\Omega }f(๐)^q(H(๐)U_q)}`$ $`=`$ $`0,`$ (43)
with $`H`$ given in Eq. (23).
The probability distribution $`f(๐)`$ reads
$$f(๐)=\frac{1}{\overline{Z}_q}[1(1q)\beta ^{}(H(๐)U_q)]^{\frac{1}{1q}},$$
(44)
where
$$\overline{Z}_q=๐\mathrm{\Omega }[1(1q)\beta ^{}(H(๐)U_q)]^{\frac{1}{1q}}.$$
(45)
In these equations, $`\beta ^{}`$ is the Lagrange multiplier associated with the constraint in Eq. (43).
We now follow the steps indicated in Ref. and summarized in Section II. Our interest lies in the interval $`0<q<1`$ again. First, define
$$R_1=\frac{\mathrm{\Gamma }\left(\frac{2q}{1q}\right)}{\mathrm{\Gamma }\left(\frac{2q}{1q}+\frac{DN}{2}\right)},$$
(46)
and
$$R_2=\left[\frac{2\pi mc}{(1q)\beta ^{}}\right]^{\frac{DN}{2}}\left[1+(1q)\frac{\beta ^{}U_q}{c}\right]^{\frac{1}{1q}+\frac{DN}{2}}.$$
(47)
The associated generalized partition function is given by
$$\overline{Z}_q(\beta ^{})=R_1\frac{V^N}{N!h^{DN}}R_2,$$
(48)
while, introducing
$$G_1=\frac{DN}{2\beta ^{}\overline{Z}_q(\beta ^{})},$$
(49)
the generalized internal energy turns out to be
$$U_q=G_1R_1\frac{V^N}{N!h^{DN}}R_2.$$
(50)
After replacing $`\overline{Z}_q(\beta ^{})`$ by Eq. (48), one finds
$$U_q=\frac{DN}{2\beta ^{}}.$$
(51)
Now, taking advantage of the fact that (Cf. Eq. (20))
$$\beta ^{}=\frac{1}{k_BT},$$
(52)
one is immediately led to
$$U_q=\frac{D}{2}Nk_BT,$$
(53)
i.e., to the ordinary result. Note that this result is obtained here without going to the limit $`q1`$. By following the methodology advanced in Ref. , we see that, with the identification of Eq. (2), Tsallis thermostatistics is able to reproduce ordinary results independently of the specific choice of the value of $`q`$.
Now, let us discuss statistical properties of the single-particle energies. Following again Section II, we focus our attention on the single particle Hamiltonians pertaining to the $`i`$th and the $`j`$th particles, $`H_i=๐ฉ_i^2/2m`$ and $`H_j=๐ฉ_j^2/2m`$, respectively. Their generalized variance, covariance and correlation coefficient are defined collectively in Eq. (yields
$`(\mathrm{\Delta }_qH_i)^2`$ $`={\displaystyle \frac{1}{2(\beta ^{})^2}}{\displaystyle \frac{2D+(1q)D^2(N1)}{42q(1q)DN}},`$ (54)
$`C_q(H_i,H_j)`$ $`={\displaystyle \frac{1}{2(\beta ^{})^2}}{\displaystyle \frac{(1q)D^2}{42q(1q)DN}},`$ (55)
$`\rho (H_i,H_j)`$ $`={\displaystyle \frac{(1q)D}{2+(1q)D(N1)}},`$ (56)
respectively. In the limit $`q1`$, one has
$`(\mathrm{\Delta }_qH_i)^2`$ $`{\displaystyle \frac{D}{2(\beta ^{})^2}},`$ (57)
$`C_q(H_i,H_j)`$ $`0,`$ (58)
$`\rho (H_i,H_j)`$ $`0.`$ (59)
We see that the limit $`q1,`$ $`C_q`$ and $`\rho `$ behave as in the Tsallis-Mendes-Plastino formalism in Ref. . However, for $`(\mathrm{\Delta }_qH_i)^2`$, we find something new: the result becomes independent of the density. It depends only on the temperature, as it does in the ordinary Boltzmann-Gibbs treatment.
## V Conclusions
In this work, we have addressed the ideal gas problem within the framework of Tsallis thermostatistics reformulated by the optimal Lagrange multipliers (OLM) method advanced in Ref. . Three interesting observations have been presented:
* The internal energy was found to be given by
$$U_q=\frac{D}{2}Nk_BT,$$
(60)
which is the same as the ordinary result obtained by Boltzmann-Gibbs theory. We derived this result without going to the limit $`q1`$. That is, it is valid for all $`q`$.
* The correlation coefficient for the particle energies become density-independent in OLM, in contrast to that in the Tsallis-Mendes-Plastino formalism in Ref. .
* The constant $`k_T`$ in the definition of the Tsallis entropy may depend on $`q`$ and, therefore, on the number of particles. In the Boltzmann-Gibbs canonical ensemble theory, the limit particle number $`N`$ going to infinity is always in mind, both explicitly and implicitly. This indicates that $`k_Tk_B`$ and $`q1`$ may correspond to $`N\mathrm{}Avogadro^{}snumber`$.
###### Acknowledgements.
The financial support of the National Research Council (CONICET) of Argentina is gratefully acknowledged. F. Pennini acknowledges financial support from UNLP, Argentina.
|
warning/0006/quant-ph0006039.html
|
ar5iv
|
text
|
# Function-dependent Phase Transform in Quantum Computing
## Abstract
We construct a quantum algorithm that performs function-dependent phase transform and requires no initialization of an ancillary register. The algorithm recovers the initial state of an ancillary register regardless of whether its state is pure or mixed. Thus we can use any qubits as an ancillary register even though they are entangled with others and are occupied by other computational process. We also show that our algorithm is optimal in the sense of the number of function evaluations.
Quantum computation is based on three quantum phenomena: superposition of states, quantum interference, and quantum entanglement. These effects enable exponential speedups in the solutions of certain problems and allow one to transgress some boundaries of classical computational complexity theory . Most known quantum algorithms rely on conditional phase transform the realization of which is accomplished by the quantum Fourier transform or the Walsh-Hadamard operator together with the unitary operator evaluating a given function or quantum oracle. In general, conditional phase transform can be described by the operation $`_{x=0}^{N1}\alpha _x|x_{x=0}^{N1}\mathrm{exp}[2\pi if(x)/M]\alpha _x|x`$ for a function $`f:_N_M`$, which we call $`f`$-dependent phase transform. The resulting interference pattern facilitates determining global property of the underlying function.
We need a quantum circuit evaluating a function to perform function-dependent phase transform and unitary evolution of quantum computational process requires an ancillary register from which we have to extract the desired relative phases conditioned on the given function. All previous quantum algorithms resort to initialization of the ancillary register before the computation. We may ask a question: Is it possible to perform function-dependent phase transform without initializing and deforming the state of the ancillary register? If it were possible, energy dissipation caused by initialization process would be avoidable and any register that could contain useful information to be preserved could temporarily be used as an ancillary register. In this work we construct a quantum algorithm that implements function-dependent phase transform without initializing an ancillary register. Furthermore, the application of the constructed algorithm retrieves the initial state of the ancillary register. Thus the ancillary register can consist of any qubits collected from any other registers even though they are being used in other computation which can proceed after carrying out their auxiliary duty in function-dependent phase transform. We show that to realize function-dependent phase transform at least two operations dependent on the given function are necessary. Thus the presented algorithm is optimal in the sense that it involves only two function evaluations. Of course, if any kind of initialization is involved, one function evaluation is sufficient.
The $`f`$-dependent phase transform $`_{k,f}:|x\omega _M^{kf(x)}|x`$ plays an important role in quantum algorithms where $`\omega _M=\mathrm{exp}(2\pi i/M)`$ is a primitive $`M`$-th root of unity and $`k_M`$ may be chosen appropriately depending on the given problems. For simplicity, we assume that $`N`$ and $`M`$ are powers of 2, that is, $`N=2^n`$ and $`M=2^m`$ for some nonnegative integers $`n`$ and $`m`$. In order for the information on the given function to be encoded in the phases it is necessary to evaluate the given function on quantum computer. On quantum computer the evaluation of a function is performed by the unitary operation $`๐ฐ_f:|x|y|x|y+f(x)`$ for $`x_N`$ and $`y_M`$. The first $`n`$-qubit register we call the control register contains the states we wish to interfere. The second $`m`$-qubit register called the function or ancillary register is used to draw relative phase changes in the first register. The superposition principle of quantum mechanics allows us to prepare the computer in a coherent superposition of input states and to compute exponentially many values of $`f`$ in superposition with a single application of $`๐ฐ_f`$. This phenomenon is the basis for quantum parallelism which leads to a completely new model of computation. In view of the second register the function evaluation adopts a translation operator $`๐ฏ_z:|y|y+z`$ where $`z`$ is dependent on the state of the first register. The operator $`๐ฐ_f`$ can be described in terms of the translation operator on the second register;
$$๐ฐ_f:\underset{x,y}{}\alpha _{xy}|x|y\underset{x,y}{}\alpha _{xy}|x๐ฏ_{f(x)}|y.$$
(1)
We note that to implement $`_{k,f}`$ at least two registers are necessary due to $`๐ฐ_f`$. By adding an ancillary register the effect of the $`_{k,f}`$ on the control register can be viewed as phase changes in the ancillary register dependent on the states of the control register. To be more specific, we define a unitary operator $`๐ฅ_{k,z}:|y\omega _M^{kz}|y`$ for $`y_M`$. Then $`_{k,f}I`$ can explicitly be written by
$$_{k,f}:\underset{x,y}{}\alpha _{xy}|x|y\underset{x,y}{}\alpha _{xy}|x๐ฅ_{k,f(x)}|y.$$
(2)
We note that $`๐ฅ_{k,z}`$ has one eigenvalue $`\omega _M^{kz}`$ and the corresponding eigenspace is the whole Hilbert space. Due to the expressions (1) and (2) we can concentrate on the operations of the ancillary register.
Especially when $`f`$ is the identity map $``$, $`_{k,}`$ maps $`|y`$ to $`\omega _M^{ky}|y`$, in which the phase-encoded information depends on its state, and can be obtained by $`^{}๐ฏ_k^{}`$ where $``$ is the quantum Fourier transform. A quantum algorithm to implement $`๐ฅ_{k,z}`$ can be realized using $`_{k,}`$ and $`๐ฏ_z`$. We prepare an arbitrary $`m`$-qubit register whose state is $`|\mathrm{\Psi }=_{y=0}^{M1}\alpha _y|y`$ and proceed the following algorithm: (i) Apply $`๐ฏ_z`$. (ii) Apply $`_{k,}`$. (iii) Apply $`๐ฏ_z^{}=๐ฏ_z`$. (iv) Apply $`_{k,}^{}=^{}๐ฏ_k`$. Then the state evolves as follows:
$`|\mathrm{\Psi }`$ $`\stackrel{๐ฏ_z}{}`$ $`{\displaystyle \underset{y=0}{\overset{M1}{}}}\alpha _y|y+z`$ (3)
$`\stackrel{_{k,}}{}`$ $`{\displaystyle \underset{y=0}{\overset{M1}{}}}\omega _M^{k(y+z)}\alpha _y|y+z`$ (4)
$`\stackrel{๐ฏ_z^{}}{}`$ $`{\displaystyle \underset{y=0}{\overset{M1}{}}}\omega _M^{k(y+z)}\alpha _y|y`$ (5)
$`\stackrel{_{k,}^{}}{}`$ $`\omega _M^{kz}|\mathrm{\Psi }.`$ (6)
Therefore we get $`_{k,}^{}๐ฏ_z^{}_{k,}๐ฏ_z|\mathrm{\Psi }=\omega _M^{kz}|\mathrm{\Psi }`$ for an arbitrary $`|\mathrm{\Psi }`$, namely, we have
$$๐ฅ_{k,z}=\omega _M^{kz}=_{k,}^{}๐ฏ_z^{}_{k,}๐ฏ_z.$$
(7)
Theorem 1. There exists a quantum algorithm to implement $`๐ฅ_{k,z}`$ using two $`๐ฏ_{\pm z}`$.
The algorithm for $`๐ฅ_{k,z}`$ is not unique. All cyclic permutations of the steps are identical. For example, we can start at Step (ii), perform successive steps, and end at Step (i). In fact, if we use the notation $`[A,B]=ABA^1B^1`$ then by Eq. (7) we have
$`๐ฅ_{k,z}`$ $`=`$ $`[_{k,}^{},๐ฏ_z^{}]=[๐ฏ_z,_{k,}^{}]`$ (8)
$`=`$ $`[_{k,},๐ฏ_z]=[๐ฏ_z^{},_{k,}]`$ (9)
with its inverse $`๐ฅ_{k,z}=[_{k,},๐ฏ_z^{}]=[๐ฏ_z,_{k,I}]=[_{k,}^{},๐ฏ_z]=[๐ฏ_z^{},_{k,}^{}]`$. Furthermore, noting that $`๐ฎ_{k,}=๐ฏ_k^{}`$ maps $`|y`$ to $`\omega _M^{ky}|y`$ one can easily check that
$$๐ฅ_{k,z}=๐ฎ_{k,}๐ฏ_z๐ฎ_{k,}๐ฏ_z,$$
(10)
which also offers another implementation. We remark that $`๐ฎ_{k,}^{}=๐ฎ_{k,}`$. Therefore there are many methods to implement $`๐ฅ_{k,z}`$. However, the number of $`๐ฏ_{\pm z}`$ in each implementation is always equal to two and cannot be reduced.
Let us suppose that there exists a quantum algorithm implementing $`๐ฅ_{k,z}`$ where the only way to implement the given information on $`z`$ is through $`๐ฏ_{\pm z}`$. Then the dependence of $`z`$ requires at least one $`๐ฏ_{\pm z}`$ at a certain step and hence the overall unitary operation performed by the algorithm can be written by $`๐ฑ_2๐ฏ_{\pm z}๐ฑ_1=\omega _M^{kz}`$ for some unitary operators $`๐ฑ_1`$ and $`๐ฑ_2`$. Since $`๐ฑ_1๐ฑ_2๐ฏ_{\pm z}=\omega _M^{kz}`$, it is enough to consider a unitary operator $`๐ฑ`$ such that $`๐ฑ๐ฏ_{\pm z}=\omega _M^{kz}`$. Since $`๐ฑ=\omega _M^{kz}๐ฏ_{\pm z}^{}`$, $`๐ฑ`$ depends on $`z`$. Thus in some another step of the algorithm we have to use information on $`z`$ once more and so the overall procedure includes at least two translations by $`\pm z`$. This observation will later be used in showing that our algorithm for $`_{k,f}`$ is optimal.
We now turn to the $`f`$-dependent phase transform $`_{k,f}`$. We let $`|\mathrm{\Phi }=_{x=0}^{N1}\alpha _x|x`$ and $`|\mathrm{\Psi }=_{y=0}^{M1}\beta _y|y`$ be the respective states of the control and the ancillary registers. It is noted that no initialization is involved during the preparation of the registers. By inspecting Eqs. (1) and (7) the algorithm (6) for $`๐ฅ_{k,z}`$ leads to an algorithm for $`_{k,f}`$: (i) Apply $`๐ฐ_f`$. (ii) Apply $`_{k,}`$. (iii) Apply $`๐ฐ_f^{}=๐ฐ_f`$. (iv) Apply $`_{k,}^{}=_{k,}`$. This procedure makes the state of the registers evolve as follows:
$`|\mathrm{\Phi }|\mathrm{\Psi }`$ $`\stackrel{๐ฐ_f}{}`$ $`{\displaystyle \underset{x=0}{\overset{N1}{}}}{\displaystyle \underset{y=0}{\overset{M1}{}}}\alpha _x\beta _y|x|y+f(x)`$ (11)
$`\stackrel{_{k,}}{}`$ $`{\displaystyle \underset{x=0}{\overset{N1}{}}}{\displaystyle \underset{y=0}{\overset{M1}{}}}\alpha _x\beta _y\omega _M^{k(y+f(x))}|x|y+f(x)`$ (12)
$`\stackrel{๐ฐ_f^{}}{}`$ $`{\displaystyle \underset{x=0}{\overset{N1}{}}}{\displaystyle \underset{y=0}{\overset{M1}{}}}\alpha _x\beta _y\omega _M^{k(y+f(x))}|x|y`$ (13)
$`\stackrel{_{k,}^{}}{}`$ $`\left({\displaystyle \underset{x=0}{\overset{N1}{}}}\omega _M^{kf(x)}\alpha _x|x\right)|\mathrm{\Psi }.`$ (14)
Now we discard the ancillary register. Then we obtain the $`f`$-dependent phase transform $`_{k,f}:|x\omega _M^{kf(x)}|x`$.
Theorem 2. There exists a quantum algorithm that implements function-dependent phase transform using two evaluations of a given function such that the ancillary register preserves its initial state.
Since the $`f`$-dependent phase transform $`_{k,f}`$ can be written in terms of $`๐ฅ_{k,z}`$ as in Eq. (2) and by Eq. (9) there are many methods to realize $`๐ฅ_{k,z}`$, we can conclude that the algorithm for $`_{k,f}`$ is also not unique.
In the procedure (14) we have assumed that an ancillary register is in a pure state. However, this is not an essential requirement. In fact, any mixed state is allowed. To be more precise, let $`A`$ be a quantum system to be used as an ancillary register and its state be described by the density operator $`\rho ^A`$. Then there exists a reference system $`R`$ such that the compound system $`AR`$ is in pure entangled state $`|\mathrm{\Psi }^{AR}`$ that gives rise to the given reduced state $`\rho ^A=\mathrm{Tr}_R(\rho ^{AR})`$ where $`\rho ^{AR}=|\mathrm{\Psi }^{AR}\mathrm{\Psi }^{AR}|`$ is called purification of $`\rho ^A`$. Using the Schmidt decomposition we can rewrite $`|\mathrm{\Psi }^{AR}`$ as $`_{y=0}^{M1}\alpha _y|y^A|\mathrm{\Psi }_y^R`$. We note that the states $`|\mathrm{\Psi }_y^R`$โs may not form the standard basis for the subsystem $`R`$ but just an orthonormal basis while the states $`|y^A`$โs form the standard basis for the subsystem $`A`$. Now applying the above algorithm to $`|\mathrm{\Phi }|\mathrm{\Psi }^{AR}`$ one can see that the final state becomes $`\left(_{k,f}|\mathrm{\Phi }\right)|\mathrm{\Psi }^{AR}`$. Thus our algorithm works whether the state of the ancillary register is pure or mixed. This implies that we can compose an ancillary register of any $`m`$ qubits which are collected out of any other registers even though they are still being used in other computational process and are possibly entangled with other qubits. The presented algorithm (14) recovers the initial state of the joint system $`AR`$ after extracting the desired relative phase changes. Thus the qubits in the temporarily composed register can be restored to their positions to continue the suspended computation.
Our algorithm requires two evaluations of $`f`$, i.e., $`๐ฐ_f`$ and $`๐ฐ_f^{}`$ \[or two $`๐ฐ_f`$ when Eq. (10) is applied\]. This is because we employ no initialization. We know that $`๐ฐ_f`$ causes translations in the ancillary register by Eq. (1) and that any quantum algorithm for $`๐ฅ_{k,z}`$ adopts at least two translations. Thus we have the following theorem.
Theorem 3. Any quantum algorithm that implements function-dependent phase transform without initialization requires at least two evaluations of a given function.
On the other hand, if the ancillary register is initializable only one evaluation of $`f`$ is sufficient. Indeed it is clear that $`|k`$ is an eigenvector of $`๐ฏ_z`$ with the corresponding eigenvalue $`\omega _M^{kz}`$. If we let $`|\mathrm{\Psi }=๐ฏ_k|0`$, then $`๐ฐ_f`$ maps $`|x|\mathrm{\Psi }`$ to $`\omega _M^{kf(x)}|x|\mathrm{\Psi }`$. The special case for $`k=1`$ was studied in .
Initialization in general sense is a process to transform the state of a quantum system to a definite pure state, which can later be rotated to $`|0`$ as usual or any other desired state by frame change. When we are to initialize the subsystem $`A`$ we cannot avoid corrupting the correlation between the subsystems $`A`$ and $`R`$, which can be measured by the quantum mutual entropy $`S(A:R)`$. If the subsystem $`A`$ is entangled with the reference system $`R`$, that is, $`S(A:R)0`$, then even when $`\rho ^A`$ is known we cannot initialize the subsystem $`A`$ by local unitary operations on $`A`$. We note that the quantum mutual entropy is invariant under local unitary operations of product form for each subsystem. If the bipartite systems $`A`$ and $`R`$ are separable, that is, $`S(A:R)=0`$, then a certain frame change on the subsystem $`A`$ effects on initialization of the ancillary register without knowing the total state $`\rho ^{AR}`$. In this sense we say that the subsystem $`A`$ is nondestructively initializable when $`\rho ^A`$ is pure and known and initialization by a local frame change on the subsystem $`A`$ is called nondestructive initialization. If nondestructive initialization is adopted then the ancillary register regains its early state and one evaluation of a function is sufficient for function-dependent phase transform.
Let us consider a more general function $`f:_N[0,1)`$. We define $`m`$-bit approximation $`\stackrel{~}{f}:_N_M`$ of $`f`$ by $`\stackrel{~}{f}(x)=_{i=1}^ma_i2^{mi}_M`$ for $`x_N`$ where $`(0.a_1a_2\mathrm{}a_m)_2=_{i=1}^ma_i2^i`$ is an $`m`$-bit binary expansion of $`f(x)`$ for $`a_i_2`$. Then the approximate $`f`$-dependent phase transform $`_{k,\stackrel{~}{f}}`$ approximates the operation $`|xe^{2\pi if(x)}|x`$ . This approximate $`f`$-dependent phase transform can be applied to the conditional $`\gamma `$-phase transform and the $`\beta `$-phase diffusion transform in .
All known quantum algorithms are based on the effect of function-dependent phase transform and the quantum Fourier transform (or the Walsh-Hadamard operator). The quantum Fourier transform enables one to find the period of a function in polynomial time and plays an essential role in Shorโs quantum polynomial-time algorithms for the integer factoring and the discrete logarithm problems which are known to be intractable on classical computer. It also enables us to construct function-dependent phase transform without initializing an ancillary register, which can immediately be applied to most known quantum algorithms.
Deutsch and Jozsa presented a simple promise problem to determine whether a Boolean function $`f:_N_2`$ is either constant or balanced and showed that it can be solved efficiently without error on quantum computer while it requires exhaustive search to solve deterministically without error in a classical setting. The key of their algorithm is the $`\pi `$-rotation of phases controlled by the query result of quantum oracle. In this problem $`M=2`$ and $`k=1`$. Then $``$ becomes the Walsh-Hadamard operator $`๐ฒ=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right)`$ and $`๐ฏ_1=๐ฏ_1`$ becomes the Pauli spin operator $`\sigma _x=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)`$ which represents a bit-flip. The operator $`_{1,}`$ is a phase-flip operator $`\sigma _z=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)`$. Thus the overall scheme for $`_{1,f}`$ is $`(\sigma _z)๐ฐ_f(\sigma _z)๐ฐ_f`$ . Grover constructed a quantum algorithm that can find a particular item in expected time $`O(\sqrt{N})`$ when an unstructured list of $`N`$ items are given . His algorithm relies on the conditional phase transform $`S_f:|x(1)^{f(x)}|x`$ and the diffusion transform $`D=๐ฒS_0๐ฒ^{}`$ where $`f`$ is the Boolean function computed by an oracle, $`S_0=S_{f_0}`$, and $`f_0(x)=\delta _{0x}`$. In this case $`S_f=_{1,f}`$ with $`M=2`$. Brassard and Hรธyer combined Simonโs ZQP algorithm and Groverโs quantum search algorithm and showed that Simonโs problem can be solved on a quantum computer in worst-case polynomial time and thus is in QP class. Thus their QP algorithm mainly depends on the conditional phase transform. Chi and Kim generalized Groverโs algorithm and showed that a quantum computer can search a database by a single query when the number of solutions is equal to or more than a quarter. Their algorithm makes use of the conditional $`\gamma `$-phase transform $`S_{f,\gamma }:|xe^{i\gamma f(x)}|x`$ and the $`\beta `$-phase diffusion transform $`D_\beta =W_lS_{l,\beta }W_l^{}`$ where $`W_l`$ is any unitary transformation satisfying $`W_l|l=\frac{1}{\sqrt{N}}_{x=0}^{N1}|x`$ and $`S_{l,\beta }=S_{f_l,\beta }`$ with $`f_l(x)=\delta _{lx}`$. In this case we can use approximate function-dependent phase transform.
All conditional phase transforms fall into the category of function-dependent phase transform. Therefore our algorithm for function-dependent phase transform is directly applicable to most quantum algorithms. We note that for general positive integers $`N`$ and $`M`$ the approximate Fourier transform in can be used in our algorithm.
In summary, we generalized conditional phase transform to function-dependent phase transform and presented a quantum algorithm that performs function-dependent phase transform and does not require any kind of initialization of an ancillary register. Our algorithm recovers the initial state of the ancillary register. Thus we can compose an ancillary register of any qubits regardless of whether they are entangled with others or being used in another computational process. Our algorithm employs two evaluations of a given function and is optimal in that any quantum algorithm that implements function-dependent phase transform without initialization requires at least two evaluations of a given function.
This work was supported by the Brain Korea 21 Project.
|
warning/0006/nucl-th0006051.html
|
ar5iv
|
text
|
# Dynamics of lowโenergy nuclear forces for electromagnetic and weak reactions with the deuteron in the NambuโJonaโLasinio model of light nuclei
## 1 Introduction
Recently we have developed the NambuโJonaโLasinio model of light nuclei , or differently the nuclear NambuโJonaโLasinio (NNJL) model, invented for the description of lowโenergy nuclear forces at the quantum field theoretic level. We have shown that the NNJL model is fully motivated by QCD . The deuteron appears in the nuclear phase of QCD as a neutronโproton collective excitation, a Cooper npโpair, caused by a phenomenological local fourโnucleon interaction. Strong lowโenergy interactions of the deuteron coupled to itself and other particles are described in terms of oneโnucleon loop exchanges. The oneโnucleon loop exchanges allow to transfer nuclear flavours from an initial to a final nuclear state by a minimal way and to take into account contributions of nucleonโloop anomalies determined completely by oneโnucleon loop diagrams. The dominance of contributions of nucleonโloop anomalies is justified in the large $`N_C`$ approach to the description of nonโperturbative QCD with $`SU(N_C)`$ gauge group at $`N_C\mathrm{}`$, where $`N_C`$ is the number of quark colours.
In this paper we apply the NNJL model to the description of lowโenergy nuclear forces for electromagnetic and weak reactions with the deuteron by example of the evaluation of the cross sections for the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ for thermal neutrons caused by the $`{}_{}{}^{1}\mathrm{S}_{0}^{}{}_{}{}^{3}\mathrm{S}_{1}^{}`$ transition (the M1โtransition), the photomagnetic disintegration of the deuteron $`\gamma `$ \+ D $``$ n + p and weak reactions of astrophysical interest: 1) the solar proton burning p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, 2) the pepโprocess p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$ and 3) the reactions of neutrino and antiโneutrino disintegration of the deuteron caused by charged $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p, $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and neutral $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n + p weak currents.
Lowโenergy electromagnetic nuclear reactions with the deuteron. It is wellโknown that the reaction of the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ for thermal neutrons plays an important role for the primordial nucleosynthesis in the Big Bang model . Indeed, the deuterons produced via the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ burn to $`{}_{}{}^{4}\mathrm{He}`$ through the reactions D + p $``$ $`{}_{}{}^{3}\mathrm{He}`$ \+ $`\gamma `$ and $`{}_{}{}^{3}\mathrm{He}`$ \+ n $``$ $`{}_{}{}^{4}\mathrm{He}`$ \+ $`\gamma `$. Therefore, the correct description of the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ is essential for the theoretical prediction of the amount of the matter in Universe.
The reaction of the photomagnetic disintegration of the deuteron $`\gamma `$ \+ D $``$ n + p is related to the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ via timeโreversal invariance of strong and electromagnetic forces. The investigation of this reaction is conceived to get an additional check of our result for the M1โcapture related to the analysis of the energy dependence of the cross section calculated in the NNJL model at energies far from threshold.
The cross section $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})`$ for the neutronโproton radiative capture has been measured for thermal neutrons at the laboratory kinetic energy $`T_\mathrm{n}=0.0252\mathrm{eV}`$ that corresponds to the laboratory velocity $`v_\mathrm{n}/c=7.34\times 10^6`$ (the absolute value is $`v_\mathrm{n}=2.2\times 10^5\mathrm{cm}/\mathrm{s}`$) :
$`\sigma (\mathrm{np}\mathrm{D}\gamma )_{\mathrm{exp}}(T_\mathrm{n})=(334.2\pm 0.5)\mathrm{mb}.`$ (1.1)
For the first time the cross section for the neutronโproton radiative capture has been calculated by Austern in the Potential model approach (PMA):
$`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=(303\pm 4)\mathrm{mb}.`$ (1.2)
The observed discrepancy about 10$`\%`$ has been then explained by Riska and Brown in terms of the contributions of the exchange currents and the $`\mathrm{\Delta }(1232)`$ resonance.
Recently the evaluation of the cross section for the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ has been carried out by using Chiral perturbation theory in the framework of the Effective Field Theory (EFT) approach \[7โ10\] formulated by Weinberg within Effective Chiral Lagrangian description of nuclear forces (see also Refs. ). The theoretical results obtained in Refs. \[7โ10\] for the cross section for the neutronโproton radiative capture are rather contradictory. Indeed, in Ref. the experimental value of the cross section for the M1โcapture has been reproduced without free parameters, the definition of which demands a fit of experimental data, with an accuracy better than 1$`\%`$, $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=(334\pm 3)\mathrm{mb}`$. In turn, in the more recent publications \[8โ10\] the predictions are not so much optimistic. Indeed, in Ref. there has been obtained the value $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=297.2\mathrm{mb}`$ which is about 11$`\%`$ less compared with the experimental one Eq.(1.1). As has been stated in Ref. for the evaluation of the correct value of the cross section for the M1โcapture within the EFT one needs to add a free parameter undefined in the approach. This parameter should be fixed from the fit of the experimental value Eq.(1.1). This program has been realized in Ref. . Then, in Refs. the neutronโproton radiative capture has been calculated for the center of mass energies of the np pair up to $`T_{\mathrm{np}}1\mathrm{MeV}`$, $`T_\mathrm{n}=2T_{\mathrm{np}}`$, by including the contribution of the E1โtransition in addition to the M1. This has added new free parameters with respect to that introduced in Refs. , where only the M1โtransition has been taken into account.
In the NNJL model we calculate the cross section for the M1โcapture both in the treeโmeson approximation and by including the contributions of chiral oneโmeson loop corrections and the $`\mathrm{\Delta }(1232)`$ resonance. For the evaluation of chiral oneโmeson loop corrections we apply Chiral perturbation theory at the quark level (CHPT)<sub>q</sub> developed within the extended NambuโJonaโLasinio (ENJL) model with a linear realization of chiral $`U(3)\times U(3)`$ symmetry.
In our consideration the $`\mathrm{\Delta }(1232)`$ resonance is the RaritaโSchwinger field $`\mathrm{\Delta }_\mu ^a(x)`$, the isotopical index $`a`$ runs over $`a=1,2,3`$, having the following free Lagrangian :
$`_{\mathrm{kin}}^\mathrm{\Delta }(x)=\overline{\mathrm{\Delta }}_\mu ^a(x)[(i\gamma ^\alpha _\alpha M_\mathrm{\Delta })g^{\mu \nu }+{\displaystyle \frac{1}{4}}\gamma ^\mu \gamma ^\beta (i\gamma ^\alpha _\alpha M_\mathrm{\Delta })\gamma _\beta \gamma ^\nu ]\mathrm{\Delta }_\nu ^a(x),`$ (1.3)
where $`M_\mathrm{\Delta }=1232\mathrm{MeV}`$ is the mass of the $`\mathrm{\Delta }`$ resonance field $`\mathrm{\Delta }_\mu ^a(x)`$. In terms of the eigenstates of the electric charge operator the fields $`\mathrm{\Delta }_\mu ^a(x)`$ are given by
$`\begin{array}{cccc}& & \mathrm{\Delta }_\mu ^1(x)=\frac{1}{\sqrt{2}}\left(\begin{array}{c}\mathrm{\Delta }_\mu ^{++}(x)\mathrm{\Delta }_\mu ^0(x)/\sqrt{3}\\ \mathrm{\Delta }_\mu ^+(x)/\sqrt{3}\mathrm{\Delta }_\mu ^{}(x)\end{array}\right),\mathrm{\Delta }_\mu ^2(x)=\frac{i}{\sqrt{2}}\left(\begin{array}{c}\mathrm{\Delta }_\mu ^{++}(x)+\mathrm{\Delta }_\mu ^0(x)/\sqrt{3}\\ \mathrm{\Delta }_\mu ^+(x)/\sqrt{3}+\mathrm{\Delta }_\mu ^{}(x)\end{array}\right),& \\ & & \mathrm{\Delta }_\mu ^3(x)=\sqrt{\frac{2}{3}}\left(\begin{array}{c}\mathrm{\Delta }_\mu ^+(x)\\ \mathrm{\Delta }_\mu ^0(x)\end{array}\right).& \end{array}`$ (1.12)
The fields $`\mathrm{\Delta }_\mu ^a(x)`$ obey the subsidiary constraints: $`^\mu \mathrm{\Delta }_\mu ^a(x)=\gamma ^\mu \mathrm{\Delta }_\mu ^a(x)=0`$ \[14โ16\]. The Green function of the free $`\mathrm{\Delta }`$โfield is determined by
$`0|\mathrm{T}(\mathrm{\Delta }_\mu (x_1)\overline{\mathrm{\Delta }}_\nu (x_2))|0=iS_{\mu \nu }(x_1x_2).`$ (1.13)
In the momentum representation $`S_{\mu \nu }(x)`$ reads \[15โ17\]:
$`S_{\mu \nu }(p)={\displaystyle \frac{1}{M_\mathrm{\Delta }\widehat{p}}}\left(g_{\mu \nu }+{\displaystyle \frac{1}{3}}\gamma _\mu \gamma _\nu +{\displaystyle \frac{1}{3}}{\displaystyle \frac{\gamma _\mu p_\nu \gamma _\nu p_\mu }{M_\mathrm{\Delta }}}+{\displaystyle \frac{2}{3}}{\displaystyle \frac{p_\mu p_\nu }{M_\mathrm{\Delta }^2}}\right).`$ (1.14)
The most general form of the $`\pi \mathrm{N}\mathrm{\Delta }`$โ interaction compatible with the requirements of chiral symmetry reads :
$`_{\pi \mathrm{N}\mathrm{\Delta }}(x)={\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{2M_\mathrm{N}}}\overline{\mathrm{\Delta }}_\omega ^a(x)\mathrm{\Theta }^{\omega \phi }N(x)_\phi \pi ^a(x)+\mathrm{h}.\mathrm{c}.=`$
$`={\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{\sqrt{6}M_\mathrm{N}}}[{\displaystyle \frac{1}{\sqrt{2}}}\overline{\mathrm{\Delta }}_\omega ^+(x)\mathrm{\Theta }^{\omega \phi }n(x)_\phi \pi ^+(x){\displaystyle \frac{1}{\sqrt{2}}}\overline{\mathrm{\Delta }}_\omega ^0(x)\mathrm{\Theta }^{\omega \phi }p(x)_\phi \pi ^{}(x)`$
$`\overline{\mathrm{\Delta }}_\omega ^+(x)\mathrm{\Theta }^{\omega \phi }p(x)_\phi \pi ^0(x)\overline{\mathrm{\Delta }}_\omega ^0(x)\mathrm{\Theta }^{\omega \phi }p(x)_\phi \pi ^0(x)+\mathrm{}],`$ (1.15)
where $`M_\mathrm{N}=M_\mathrm{n}=M_\mathrm{p}=940\mathrm{MeV}`$ is the nucleon mass. The nucleon field $`N(x)`$ is the isotopical doublet with the components $`N(x)=(p(x),n(x))`$, and $`\pi ^a(x)`$ is the pion field with the components $`\pi ^1(x)=(\pi ^{}(x)+\pi ^+(x))/\sqrt{2}`$, $`\pi ^2(x)=(\pi ^{}(x)\pi ^+(x))/i\sqrt{2}`$ and $`\pi ^3(x)=\pi ^0(x)`$. The tensor $`\mathrm{\Theta }^{\omega \phi }`$ is given in Ref. : $`\mathrm{\Theta }^{\omega \phi }=g^{\omega \phi }(Z+1/2)\gamma ^\omega \gamma ^\phi `$, where the parameter $`Z`$ is arbitrary. There is no consensus on the exact value of $`Z`$. From theoretical point of view $`Z=1/2`$ is preferred . Phenomenological studies give only the bound $`|Z|1/2`$ . The value of the coupling constant $`g_{\pi \mathrm{N}\mathrm{\Delta }}`$ relative to the coupling constant $`g_{\pi \mathrm{NN}}`$ is $`g_{\pi \mathrm{N}\mathrm{\Delta }}=2g_{\pi \mathrm{NN}}`$ .
Assuming that the transition $`\mathrm{\Delta }\mathrm{N}+\gamma `$ is primarily a magnetic one the effective Lagrangian describing the $`\mathrm{\Delta }\mathrm{N}+\gamma `$ decays can be determined as \[19โ21\]:
$`_{\gamma \mathrm{N}\mathrm{\Delta }}(x)=ie{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{2M_\mathrm{N}}}\overline{N}(x)\gamma _\alpha \gamma ^5\mathrm{\Delta }_\beta ^3(x)F^{\beta \alpha }(x)+\mathrm{h}.\mathrm{c}.=`$ (1.16)
$`={\displaystyle \frac{ie}{\sqrt{6}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{M_\mathrm{N}}}[\overline{p}(x)\gamma _\alpha \gamma ^5\mathrm{\Delta }_\beta ^+(x)+\overline{n}(x)\gamma _\alpha \gamma ^5\mathrm{\Delta }_\beta ^0(x)]F^{\beta \alpha }(x)+\mathrm{h}.\mathrm{c}.,`$
where $`F^{\alpha \beta }(x)=^\alpha A^\beta (x)^\beta A^\alpha (x)`$ is the electromagnetic strength field tensor and $`A^\alpha (x)`$ is a vector potential of the electromagnetic field. The value of the coupling constant $`g_{\gamma \mathrm{N}\mathrm{\Delta }}`$ relative to the coupling constant $`g_{\pi \mathrm{NN}}`$ is $`g_{\gamma \mathrm{N}\mathrm{\Delta }}=0.14g_{\pi \mathrm{NN}}`$ caused by the $`SU(6)`$ symmetry of strong lowโenergy interactions .
The NNJL model realizes the Lagrange approach to the description of lowโenergy nuclear forces . For the evaluation of the effective Lagrangian of the transition n + p $``$ D + $`\gamma `$ it is necessary, first, to determine the effective Lagrangian of the strong lowโenergy transition n + p $``$ n + p, or more generally N + N $``$ N + N, where N = (p, n) is a nucleon field. Since the NNJL model describes lowโenergy interactions of the deuteron in terms of oneโnucleon loop exchanges the effective Lagrangian of the transition N + N $``$ N + N plays an important role. Due to the transition n + p $``$ n + p the np pair onโmass shell in the initial state transfers itself into the np pair offโmass shell couples to the deuteron and the photon through oneโnucleon loop exchanges. Then, the oneโnucleon loop diagrams are calculated at leading order in the $`1/M_\mathrm{N}`$ expansion that corresponds to the large $`N_C`$ expansion due to the proportionality $`M_\mathrm{N}N_C`$ valid in the multiโcolour QCD with $`SU(N_C)`$ gauge group at $`N_C\mathrm{}`$ .
Such a procedure of the evaluation of effective Lagrangians in the NNJL model resembles that has being used in the ENJL model \[13,23โ25\] for the derivation of effective chiral Lagrangians up to the formal replacement $`q(\overline{q})N(\overline{N})`$, where $`q(\overline{q})`$ is a quark (antiโquark) field. In the ENJL model the dominance of the leading order contributions in the $`1/M_q`$ expansion, where $`M_q`$ is a constituent quark mass, has been explained by a dynamics of quark confinement, whereas in the NNJL model the dominance of the leading contributions in the $`1/M_\mathrm{N}`$ expansion is justified by the large $`N_C`$ approach to nonโperturbative QCD.
Since relative momenta of the np pair in the reaction of the M1โcapture are smaller compared with the mass of the pion $`M_\pi =135\mathrm{MeV}`$, for the derivation of the effective Lagrangian of the strong lowโenergy transition n + p $``$ n + p we can follow the ideology of the EFT \[7โ12\] and integrate out pion degrees of freedom as well as other heavier degrees of freedom. The result of the integration can be represented in the form of two contributions, where the first is the explicit oneโpion exchange, whereas the second is a phenomenological one describing a collective contribution of both manyโpion exchanges and heavier meson degrees of freedom such as $`\sigma (660)`$, $`\rho (770)`$, $`\omega (782)`$ and so on.
The effective Lagrangian $`_{\mathrm{one}\mathrm{pion}}^{\mathrm{np}\mathrm{np}}(x)`$ of the strong lowโenergy transition n + p $``$ n + p in the oneโpion exchange approximation is defined by
$`_{\mathrm{one}\mathrm{pion}}^{\mathrm{np}\mathrm{np}}(x)={\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{M_\pi ^2}}\{[\overline{n}(x)\gamma ^5n(x)][\overline{p}(x)\gamma ^5p(x)]\mathrm{\hspace{0.17em}2}[\overline{p}(x)\gamma ^5n(x)][\overline{n}(x)\gamma ^5p(x)]\},`$ (1.17)
where $`g_{\pi \mathrm{NN}}=13.4`$ is the $`\pi \mathrm{NN}`$ coupling constant, $`n(x)`$ and $`p(x)`$ are the operators of the interpolating fields of the neutron and the proton. The effective Lagrangian Eq.(1.17) is local, as we have neglected the squared momentum transfer $`q^2`$ with respect to the squared pion mass, $`q^2M_\pi ^2`$. Since in the reaction n + p $``$ D + $`\gamma `$ the np pair couples to the deuteron and the photon in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state, we should rearrange the operators of the neutron and the proton interpolating fields in the effective interaction Eq.(1.17) by such a way to introduce the products of the np operators creating the np states with definite total spins. This rearrangement can be carried out by means of the Fierz transformation that gives
$`\gamma ^5\gamma ^5`$ $`=`$ $`{\displaystyle \frac{1}{4}}CC+{\displaystyle \frac{1}{4}}\gamma ^5CC\gamma ^5+{\displaystyle \frac{1}{4}}\gamma ^\mu CC\gamma _\mu +{\displaystyle \frac{1}{4}}\gamma ^\mu \gamma ^5CC\gamma _\mu \gamma ^5`$ (1.18)
$`+{\displaystyle \frac{1}{8}}\sigma ^{\mu \nu }CC\sigma _{\mu \nu },`$
where $`\sigma ^{\mu \nu }=(\gamma ^\mu \gamma ^\nu \gamma ^\nu \gamma ^\mu )/2`$. By virtue of Eq.(1.18) we recast the effective Lagrangian Eq.(1.17) into the form
$`_{\mathrm{one}\mathrm{pion}}^{\mathrm{np}\mathrm{np}}(x)`$ $`=`$ $`{\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{4M_\pi ^2}}\{[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]+[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5p(x)]`$ (1.19)
$`+[\overline{p}(x)n^c(x)][\overline{n^c}(x)p(x)]+3[\overline{p}(x)\gamma ^\mu n^c(x)][\overline{n^c}(x)\gamma _\mu p(x)]`$
$`+{\displaystyle \frac{3}{2}}[\overline{p}(x)\sigma ^{\mu \nu }n^c(x)][\overline{n^c}(x)\sigma _{\mu \nu }p(x)]\}.`$
Here $`\overline{n^c}(x)=n^T(x)C`$ and $`n^c(x)=C\overline{n}^T(x)`$, where $`C`$ is a charge conjugation matrix and $`T`$ is a transposition. The first two terms in the effective Lagrangian Eq.(1.19) describe the strong lowโenergy $`{}_{}{}^{1}\mathrm{S}_{0}^{}{}_{}{}^{1}\mathrm{S}_{0}^{}`$ transition of the np pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state. Thereby, the effective Lagrangian providing the $`{}_{}{}^{1}\mathrm{S}_{0}^{}{}_{}{}^{1}\mathrm{S}_{0}^{}`$ transition of the np pair and caused by the oneโpion exchange we would use in the form
$`_{\mathrm{one}\mathrm{pion}}^{\mathrm{np}\mathrm{np}}(x)=`$
$`={\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{4M_\pi ^2}}\{[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]+[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5p(x)]\}.`$ (1.20)
The phenomenological part of the effective Lagrangian responsible for the strong lowโenergy $`{}_{}{}^{1}\mathrm{S}_{0}^{}{}_{}{}^{1}\mathrm{S}_{0}^{}`$ transition of the np pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state we would choose by following the EFT ideology \[7โ12\] as well and write
$`_{\mathrm{ph}.}^{\mathrm{np}\mathrm{np}}(x)=`$
$`={\displaystyle \frac{2\pi a_{\mathrm{np}}}{M_\mathrm{N}}}\{[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]+[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5p(x)]\},`$ (1.21)
where $`a_{\mathrm{np}}=(23.75\pm 0.01)\mathrm{fm}`$ is the Sโwave scattering length of the elastic np scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state .
The appearance of the Sโwave scattering length for the definition of the phenomenological coupling constant of the strong lowโenergy transition n + p $``$ n + p, or differently lowโenergy elastic np scattering is rather natural in the EFT. Indeed, in the EFT pions are treated perturbatively and at leading order in the pionโexchange approximation for the description of the Sโwave scattering length of lowโenergy elastic np scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state Weinberg has introduced a phenomenological fourโnucleon interaction with a coupling constant
$`C_0={\displaystyle \frac{4\pi a_{\mathrm{np}}}{M_\mathrm{N}}}.`$ (1.22)
As has been stated in the EFT this constant is a collective contribution coming from the integration over all meson degrees of freedom heavier than the pionic ones. Since in our approach we separate the pionic degrees of freedom from that with masses heavier than the pion as well, the integration over these heavy meson degrees of freedom should have the same form as it is postulated in the EFT. We have only halved the coupling constant $`C_0`$ defined by Eq.(1.22) in order to distribute symmetrically phenomenological contributions between couplings $`[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]`$ and $`[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5p(x)]`$.
The total effective Lagrangian describing the strong lowโenergy transition n + p $``$ n + p of the np pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state is then determined by
$`_{\mathrm{eff}}^{\mathrm{np}\mathrm{np}}(x)=_{\mathrm{one}\mathrm{pion}}^{\mathrm{np}\mathrm{np}}(x)+_{\mathrm{ph}.}^{\mathrm{np}\mathrm{np}}(x)=`$
$`=C_{\mathrm{NN}}\{[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]+[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5p(x)]\},`$ (1.23)
where the effective coupling constant $`C_{\mathrm{NN}}`$ of the strong lowโenergy transition n + p $``$ n + p is equal to
$`C_{\mathrm{NN}}={\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{4M_\pi ^2}}{\displaystyle \frac{2\pi a_{\mathrm{np}}}{M_\mathrm{N}}}=3.27\times 10^3\mathrm{MeV}^2.`$ (1.24)
Note that the contribution of the phenomenological part to the effective coupling constant $`C_{\mathrm{NN}}`$ makes up less than 33$`\%`$.
Since nuclear forces are isotopically invariant , the effective Lagrangian of the strong lowโenergy N + N $``$ N + N transition of the NN pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state can be defined as follows
$`_{\mathrm{eff}}^{\mathrm{NN}\mathrm{NN}}(x)=C_{\mathrm{NN}}`$
$`\{[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]+[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5p(x)]`$
$`+{\displaystyle \frac{1}{2}}[\overline{n}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5n(x)]+[\overline{n}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5n(x)]`$
$`+{\displaystyle \frac{1}{2}}[\overline{p}(x)\gamma ^5p^c(x)][\overline{p^c}(x)\gamma ^5p(x)]+[\overline{p}(x)\gamma ^\mu \gamma ^5p^c(x)][\overline{p^c}(x)\gamma _\mu \gamma ^5p(x)]\}.`$ (1.25)
We would like to emphasize the effective Lagrangian Eq.(1.24) describing strong lowโenergy N + N $``$ N + N transitions is obtained in complete agreement with the EFT ideology.
In the lowโenergy limit the effective local fourโnucleon interaction Eq. (1) vanishes due to the reduction
$`[\overline{N}(x)\gamma _\mu \gamma ^5N^c(x)][\overline{N^c}(x)\gamma ^\mu \gamma ^5N(x)][\overline{N}(x)\gamma ^5N^c(x)][\overline{N^c}(x)\gamma ^5N(x)],`$ (1.26)
where $`N(x)`$ is an operator of the neutron or the proton interpolating field. Such a vanishing of the oneโpion exchange contribution to the NN potential is wellโknown in the EFT approach and the PMA . In power counting the interaction induced by the oneโpion exchange is of order $`O(p^2)`$, where $`p`$ is a relative momentum of the NN system. The former is due to the Dirac matrix $`\gamma ^5`$ which leads to the interaction between small components of Dirac bispinors of nucleon wave functions.
In the oneโnucleon loop exchange approach the contributions of the interactions $`[\overline{N}(x)\gamma _\mu \gamma ^5N^c(x)][\overline{N^c}(x)\gamma ^\mu \gamma ^5N(x)]`$ and $`[\overline{N}(x)\gamma ^5N^c(x)][\overline{N^c}(x)\gamma ^5N(x)]`$ to the amplitudes of nuclear reactions are different and do not cancel each other in the lowโenergy limit due to the dominance of nucleonโloop anomalies . This provides the interaction between large components of Dirac bispinors of nucleon wave functions that distinguishes the NNJL model from the EFT.
Lowโenergy weak nuclear reactions with the deuteron. The weak nuclear reaction p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, the solar proton burning, plays an important role in Astrophysics . It gives start for the pโp chain of nucleosynthesis in the Sun and mainโsequence stars . In the Standard Solar Model (SSM) the total (or bolometric) luminosity of the Sun $`L_{}=(3.846\pm 0.008)\times 10^{26}\mathrm{W}`$ is normalized to the astrophysical factor $`S_{\mathrm{pp}}(0)`$ for the solar proton burning. The recommended value $`S_{\mathrm{pp}}(0)=4.00\times 10^{25}\mathrm{MeVb}`$ has been found by averaging over the results obtained in the Potential model approach (PMA) and the Effective Field Theory (EFT) approach . As has been shown recently in Ref. the inverse and forward helioseismic approach confirm the recommended value of $`S_{\mathrm{pp}}(0)`$ within experimental errors on the helioseismic data and solar neutrino fluxes.
In this paper we apply the NNJL model to the description of lowโenergy nuclear forces for weak nuclear reactions with the deuteron of astrophysical interest: 1) the solar proton burning p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, 2) the pepโprocess p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$ and 3) the reactions of neutrino and antiโneutrino disintegration of the deuteron caused by charged $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p, $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and neutral $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n + p weak currents. The reactions $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p and $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p caused by charged and neutral weak currents, respectively, and induced by solar neutrinos are planned to be measured for solar neutrino experiments at Sudbury Neutrino Observatory (SNO) . In turn, the cross sections for the reactions $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p caused by charged and neutral weak currents, respectively, and induced by reactor antiโneutrinos have been recently measured by the Reinesโs experimental group . In the sense of charge independence of weak interaction strength the observation of the reaction $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n is equivalent to the observation of the reaction of the solar proton burning p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ in the terrestrial laboratories.
The paper is organized as follows. In Section 2 we evaluate the amplitude of the M1โcapture n + p $``$ D + $`\gamma `$ in the treeโmeson approximation. The contribution of lowโenergy elastic np scattering to the amplitude of the process n + p $``$ D + $`\gamma `$ is obtained in agreement with lowโenergy nuclear phenomenology. In Section 3 we evaluate contributions of chiral oneโmeson loop corrections to the amplitude of the M1โcapture in Chiral perturbation theory at the quark level (CHPT)<sub>q</sub> developed within the ENJL model with a linear realization of chiral $`U(3)\times U(3)`$ symmetry. In Section 4 we include the contribution of the $`\mathrm{\Delta }(1232)`$ resonance and analyse the total cross section for the neutronโproton radiative capture for thermal neutrons and compare it with experimental data. In Section 5 we treat the reaction of the photomagnetic disintegration of the deuteron $`\gamma `$ \+ D $``$ n + p related to the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ via timeโreversal invariance and analyse the energy dependence of the cross section at energies far from threshold. In Sect. 6 we evaluate the amplitude of the solar proton burning. We show that the contribution of lowโenergy elastic pp scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state with the Coulomb repulsion is described in agreement with lowโenergy nuclear phenomenology in terms of the Sโwave scattering length and the effective range. In Sect. 7 we evaluate the astrophysical factor for the solar proton burning and obtain the value $`S_{\mathrm{pp}}(0)=4.08\times 10^{25}\mathrm{MeV}\mathrm{b}`$ agreeing good with the recommended one $`S_{\mathrm{pp}}(0)=4.00\times 10^{25}\mathrm{MeV}\mathrm{b}`$. In Sect. 8 we evaluate the cross section for the neutrino disintegration of the deuteron $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p with respect to $`S_{\mathrm{pp}}(0)`$. In Sect. 9 we adduce the evaluation of the astrophysical factor $`S_{\mathrm{pep}}(0)`$ for the pepโprocess relative to $`S_{\mathrm{pp}}(0)`$. In Sects. 10 and 11 we evaluate the cross sections for the antiโneutrino disintegration of the deuteron $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p and average them over the antiโneutrino energy spectrum. The average values of the cross sections agree well with experimental data . The cross sections for the weak nuclear reactions of astrophysical interest are calculated at zero contribution of the nucleon tensor current . This makes the description of lowโenergy nuclear forces within the NNJL model compatible with the predictions of the SSM . In more detail this is discussed the Conclusion and Appendix. In the Conclusion we discuss the obtained results. In Appendix we evaluate the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ of the lowโenergy weak transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$. The contribution of the nucleon tensor current to the effective Lagrangians of lowโenergy weak nuclear transitions of astrophysical interest like p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ and so on is evaluated and discussed.
## 2 The M1โcapture in the treeโmeson approximation
Since the NNJL model realizes a Lagrange approach to the description of lowโenergy nuclear forces , first thing what we have to do is to evaluate the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{D}\gamma }(x)`$ of the transition n + p $``$ D + $`\gamma `$. In the treeโmeson approximation the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{D}\gamma }(x)`$ is defined by oneโnucleon loop diagrams depicted in Figs. 1 and 2. The evaluation of these diagrams at leading order in the large $`N_C`$ expansion yields
$`_{\mathrm{eff}}^{\mathrm{np}\mathrm{D}\gamma }(x)`$ $`=`$ $`(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{e}{2M_\mathrm{N}}}{\displaystyle \frac{g_\mathrm{V}}{4\pi ^2}}C_{\mathrm{NN}}D_{\mu \nu }^{}(x){}_{}{}^{}F_{}^{\mu \nu }(x)[\overline{p^c}(x)\gamma ^5n(x)]`$ (2.1)
$`+`$ $`i(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{e}{2M_\mathrm{N}}}{\displaystyle \frac{g_\mathrm{V}}{4\pi ^2}}C_{\mathrm{NN}}M_\mathrm{N}D_\mu ^{}(x){}_{}{}^{}F_{}^{\mu \nu }(x)[\overline{p^c}(x)\gamma _\nu \gamma ^5n(x)],`$
where $`{}_{}{}^{}F_{}^{\mu \nu }(x)=\frac{1}{2}\epsilon ^{\mu \nu \alpha \beta }F_{\alpha \beta }(x)`$, $`\mu _\mathrm{p}=2.793`$ and $`\mu _\mathrm{n}=\mathrm{\hspace{0.17em}1.913}`$ are the magnetic dipole moments of the proton and the neutron measured in nuclear magnetons, $`D_\mu (x)`$ is the operator of the interpolating field of the deuteron and $`D_{\mu \nu }(x)=_\mu D_\nu (x)_\nu D_\mu (x)`$, $`g_\mathrm{V}`$ is a phenomenological coupling constant of the NNJL model related to the electric quadrupole moment of the deuteron $`Q_\mathrm{D}=0.286\mathrm{fm}^2`$ : $`g_\mathrm{V}^2=2\pi ^2Q_\mathrm{D}M_\mathrm{N}^2`$.
The matrix element of the transition n + p $``$ D + $`\gamma `$ we define by a usual way
$`{\displaystyle d^4xD(k_\mathrm{D})\gamma (k)|_{\mathrm{eff}}^{\mathrm{np}\mathrm{D}\gamma }(x)|n(p_1)p(p_2)}=`$
$`=(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+kp_1p_2){\displaystyle \frac{(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )}{\sqrt{2E_1V\mathrm{\hspace{0.17em}2}E_2V\mathrm{\hspace{0.17em}2}E_\mathrm{D}V\mathrm{\hspace{0.17em}2}\omega V}}},`$ (2.2)
where $`E_i(i=1,2,\mathrm{D})`$ and $`\omega `$ are the energies of the neutron, the proton, the deuteron and the photon, $`V`$ is the normalization volume.
The wave functions of the initial $`|n(p_1)p(p_2)`$ and final $`D(k_\mathrm{D})\gamma (k)|`$ state we take in the usual form
$`|n(p_1)p(p_2)`$ $`=`$ $`a_\mathrm{n}^{}(\stackrel{}{p}_1,\sigma _1)a_\mathrm{p}^{}(\stackrel{}{p}_2,\sigma _2)|0,`$
$`D(k_\mathrm{D})\gamma (k)|`$ $`=`$ $`0|a_\mathrm{D}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})a(\stackrel{}{k},\lambda ),`$ (2.3)
where $`a_\mathrm{n}^{}(\stackrel{}{p}_1,\sigma _1)`$ and $`a_\mathrm{p}^{}(\stackrel{}{p}_2,\sigma _2)`$ are the operators of creation of the neutron and the proton, and $`a_\mathrm{D}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})`$ and $`a(\stackrel{}{k},\lambda )`$ are the operators of annihilation of the deuteron and the photon.
The matrix element of the transition n + p $``$ D + $`\gamma `$ reads
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )`$ $`=`$ $`(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{e}{2M_\mathrm{N}}}{\displaystyle \frac{g_\mathrm{V}}{4\pi ^2}}C_{\mathrm{NN}}\epsilon ^{\alpha \beta \mu \nu }k_\alpha e_\beta ^{}(k,\lambda )e_\mu ^{}(k_\mathrm{D},\lambda _\mathrm{D})`$ (2.4)
$`\times [\overline{u^c}(p_2)(2k_{\mathrm{D}\nu }M_\mathrm{N}\gamma _\nu )\gamma ^5u(p_1)],`$
where $`e_\beta ^{}(k,\lambda )`$ and $`e_\mu ^{}(k_\mathrm{D},\lambda _\mathrm{D})`$ are the 4โvectors of the polarization of the photon and the deuteron, then $`\overline{u^c}(p_2)`$ and $`u(p_1)`$ are the bispinorial wave functions of the proton and the neutron, respectively, normalized by $`\overline{u^c}(p_2)u^c(p_2)=\mathrm{\hspace{0.17em}2}M_\mathrm{N}`$ and $`\overline{u}(p_1)u(p_1)=2M_\mathrm{N}`$.
In the lowโenergy limit when
$`[\overline{u^c}(p_2)\gamma _\nu \gamma ^5u(p_1)]g_{\nu 0}[\overline{u^c}(p_2)\gamma ^5u(p_1)]`$ (2.5)
and $`k_{\mathrm{D}\nu }g_{\nu 0}\mathrm{\hspace{0.17em}2}M_\mathrm{N}`$ the matrix element Eq.(2.4) acquires the form
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )`$ $`=`$ $`e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})`$ (2.6)
$`\times [\overline{u^c}(p_2)\gamma ^5u(p_1)].`$
The evaluation of the matrix element Eq.(2.6), the effective vertex of the n + p $``$ D + $`\gamma `$ transition, we have carried out with the wave functions of the neutron and the proton in the form of the plane waves. However, a physical $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state of the np pair is defined by lowโenergy nuclear forces. For the description of the contribution of lowโenergy nuclear forces to the physical $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state of the np pair coupled to the deuteron and the photon we suggest to sum an infinite series of oneโnucleon bubbles with vertices defined by the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{np}}(x)`$ Eq.(1). The result of the summation can be represented in the following form
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )=e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})[\overline{u^c}(p_2)\gamma ^5u(p_1)]`$
$`\times {\displaystyle \frac{1}{1+{\displaystyle \frac{C_{\mathrm{NN}}}{16\pi ^2}}{\displaystyle \frac{d^4q}{\pi ^2i}\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}\widehat{P}\widehat{Q}}\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}\widehat{Q}}\right\}}}},`$ (2.7)
where $`P=p_1+p_2=(2\sqrt{p^2+M_\mathrm{N}^2},\stackrel{}{0})`$ is the 4โmomentum of the np pair in the center of mass frame. Then, $`Q=aP+bK=a(p_1+p_2)+b(p_1p_2)`$ is an arbitrary shift of virtual momentum with arbitrary parameters $`a`$ and $`b`$, and in the center of mass frame $`K=p_1p_2=(0,2\stackrel{}{p})`$. The explicit dependence of the momentum integral on $`Q`$ can be evaluated by means of the GertseinโJackiw procedure (see also Ref. ). It is given by
$`{\displaystyle \frac{d^4q}{\pi ^2i}\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}\widehat{P}\widehat{Q}}\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}\widehat{Q}}\right\}}=`$
$`={\displaystyle \frac{d^4q}{\pi ^2i}\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}\widehat{P}}\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}}\right\}}2a(a+1)P^22b^2K^2.`$ (2.8)
For the evaluation of the momentum integral over $`q`$ we would keep only the leading order contributions in the $`1/M_\mathrm{N}`$ expansion caused by the large $`N_C`$ expansion . This yields
$`{\displaystyle \frac{d^4q}{\pi ^2i}\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}\widehat{P}\widehat{Q}}\gamma ^5\frac{1}{M_\mathrm{N}\widehat{q}\widehat{Q}}\right\}}=`$
$`=8a(a+1)M_\mathrm{N}^2+8(b^2a(a+1))p^2i\mathrm{\hspace{0.17em}8}\pi M_\mathrm{N}p.`$ (2.9)
The expression Eq.(2) we reduce to the form
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )`$ $`=`$ $`e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})[\overline{u^c}(p_2)\gamma ^5u(p_1)]`$ (2.10)
$`\times {\displaystyle \frac{๐ต}{1{\displaystyle \frac{1}{2}}r_{\mathrm{np}}a_{\mathrm{np}}p^2+ia_{\mathrm{np}}p}}.`$
Here we have denoted
$`a_{\mathrm{np}}`$ $`=`$ $`{\displaystyle \frac{C_{\mathrm{NN}}M_\mathrm{N}}{2\pi }}๐ต,r_{\mathrm{np}}=(b^2a(a+1)){\displaystyle \frac{2}{\pi }}{\displaystyle \frac{1}{M_\mathrm{N}}},`$
$`{\displaystyle \frac{1}{๐ต}}`$ $`=`$ $`1{\displaystyle \frac{a(a+1)}{2\pi ^2}}C_{\mathrm{NN}}M_\mathrm{N}^2,`$ (2.11)
where $`r_{\mathrm{np}}=2.75\pm 0.05\mathrm{fm}`$ is the effective range of lowโenergy elastic np scattering .
Renormalizing the wave functions of nucleons $`\sqrt{๐ต}u(p_1)u(p_1)`$ and $`\sqrt{๐ต}u(p_2)u(p_2)`$ we arrive at the expression
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )`$ $`=`$ $`e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})[\overline{u^c}(p_2)\gamma ^5u(p_1)]`$ (2.12)
$`\times {\displaystyle \frac{1}{1{\displaystyle \frac{1}{2}}r_{\mathrm{np}}a_{\mathrm{np}}p^2+ia_{\mathrm{np}}p}},`$
where the factor $`1/(1\frac{1}{2}r_{\mathrm{np}}a_{\mathrm{np}}p^2+ia_{\mathrm{np}}p)`$ describes the contribution of lowโenergy nuclear forces to the physical $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state of the np pair coupled to the deuteron and the photon. It has the form of the amplitude of lowโenergy elastic np scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state in complete agreement with lowโenergy nuclear phenomenology . By using the relation expressing the phase shift $`\delta _{\mathrm{np}}(p)`$ of lowโenergy elastic np scattering in terms of the Sโwave scattering length $`a_{\mathrm{np}}`$ and the effective range $`r_{\mathrm{np}}`$
$`\mathrm{ctg}\delta _{\mathrm{np}}(p)={\displaystyle \frac{1}{a_{\mathrm{np}}}}+{\displaystyle \frac{1}{2}}r_{\mathrm{np}}p^2`$ (2.13)
we can recast the factor $`1/(1\frac{1}{2}r_{\mathrm{np}}a_{\mathrm{np}}p^2+ia_{\mathrm{np}}p)`$ into the form
$`{\displaystyle \frac{1}{1{\displaystyle \frac{1}{2}}r_{\mathrm{np}}a_{\mathrm{np}}p^2+ia_{\mathrm{np}}p}}=e^{i\delta _{\mathrm{np}}\left(p\right)}{\displaystyle \frac{\mathrm{sin}\delta _{\mathrm{np}}(p)}{a_{\mathrm{np}}p}}.`$ (2.14)
In terms of the phase shift $`\delta _{\mathrm{np}}(p)`$ the expression Eq.(2.12) reads
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )`$ $`=`$ $`e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})`$ (2.15)
$`\times [\overline{u^c}(p_2)\gamma ^5u(p_1)]e^{i\delta _{\mathrm{np}}\left(p\right)}{\displaystyle \frac{\mathrm{sin}\delta _{\mathrm{np}}(p)}{a_{\mathrm{np}}p}}.`$
In the NNJL model lowโenergy nuclear forces between the neutron and the proton in the physical deuteron state are described by the oneโnucleon loop exchanges in terms of the phenomenological coupling constant $`g_\mathrm{V}`$ which is defined by the electric quadrupole moment of the deuteron $`Q_\mathrm{D}`$, $`g_\mathrm{V}^2=2\pi ^2Q_\mathrm{D}M_\mathrm{N}^2`$ . The electric quadrupole moment of the deuteron is caused by nuclear tensor forces . Therefore, the relation $`g_\mathrm{V}^2=2\pi ^2Q_\mathrm{D}M_\mathrm{N}^2`$ confirms at the quantum field theoretic level the fact pointed out by Blatt and Weisskopf that the existence of a bound triplet state of the neutronโproton system would be entirely due to the tensor force . Thus, in NNJL model through the phenomenological coupling constant $`g_\mathrm{V}`$ tensor forces govern the existence of the deuteron as a bound neutronโproton triplet spin state and strength of lowโenergy interactions of the deuteron with nucleons and other particles in terms of the oneโnucleon loop exchanges. The evaluation of oneโnucleon loop diagrams at leading order in the $`1/M_\mathrm{N}`$ expansion, or in the large $`N_C`$ expansion, reduces a momentum dependence of nucleon diagrams to the trivial form accounting for only the Lorentz covariant properties of the interaction. In this approach the deuteron looks like a pointโlike particle. Such a representation is enough for the evaluation of effective Lagrangians of different lowโenergy nuclear transitions with the deuteron in an initial or a final state describing effective vertices of lowโenergy nuclear transitions defined at their thresholds. However, for the evaluation of amplitudes of lowโenergy nuclear reactions for energies far from threshold one needs to take into account a spatial smearing of the physical deuteron caused by a finite radius $`r_\mathrm{D}=1/\sqrt{\epsilon _\mathrm{D}M_\mathrm{N}}=4.319\mathrm{MeV}`$ determined by the binding energy of the deuteron $`\epsilon _\mathrm{D}=2.225\mathrm{MeV}`$. The spatial smearing of the physical deuteron we introduce phenomenologically in the form
$`F_\mathrm{D}(p)={\displaystyle \frac{1}{1+r_\mathrm{D}^2p^2}},`$ (2.16)
that is nothing more than the momentum representation of the approximate $`{}_{}{}^{3}\mathrm{S}_{1}^{}`$ wave state of the deuteron .
Substituting Eq.(2.16) in Eq.(2.15) we obtain the amplitude of the M1โcapture calculated in the NNJL model:
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )`$ $`=`$ $`e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})`$ (2.17)
$`\times [\overline{u^c}(p_2)\gamma ^5u(p_1)]e^{i\delta _{\mathrm{np}}\left(p\right)}{\displaystyle \frac{\mathrm{sin}\delta _{\mathrm{np}}(p)}{a_{\mathrm{np}}p}}{\displaystyle \frac{1}{1+r_\mathrm{D}^2p^2}}.`$
For thermal neutrons the kinetic energy of the relative movement of the np pair is of order $`T_{\mathrm{np}}10^8\mathrm{MeV}`$ . This yields the relative momentum of the np pair to be smaller compared with the binding energy of the deuteron, $`p3\times 10^3\mathrm{MeV}\epsilon _\mathrm{D}=2.225\mathrm{MeV}`$. Therefore, for thermal neutrons without loss of generality we can calculate the amplitude of the M1โcapture setting $`p=0`$:
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )`$ $`=`$ $`e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})`$ (2.18)
$`\times [\overline{u^c}(p_2)\gamma ^5u(p_1)].`$
Thus, we have obtained that the amplitude of the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ for thermal neutrons coincides with the matrix element of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{D}\gamma }(x)`$ given by Eq.(2.6).
The cross section for the M1โcapture for thermal neutrons calculated in the treeโmeson approximation is then defined by
$`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})={\displaystyle \frac{1}{v_\mathrm{n}}}(\mu _\mathrm{p}\mu _\mathrm{n})^2{\displaystyle \frac{25}{64}}{\displaystyle \frac{\alpha }{\pi ^2}}Q_\mathrm{D}C_{\mathrm{NN}}^2M_\mathrm{N}\epsilon _\mathrm{D}^3\left(1+{\displaystyle \frac{1}{2}}{\displaystyle \frac{T_\mathrm{n}}{\epsilon _\mathrm{D}}}\right)^3=276\mathrm{mb}.`$ (2.19)
The theoretical value $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=276\mathrm{mb}`$ is about 17$`\%`$ smaller compared with the experimental one $`\sigma (\mathrm{np}\mathrm{D}\gamma )_{\mathrm{exp}}(T_\mathrm{n})=(334.2\pm 0.5)\mathrm{mb}`$. Thus, in the treeโmeson approximation the NNJL model predicts the cross section for the M1โcapture for thermal neutrons somewhat worse than the EFT, $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=297.2\mathrm{mb}`$ . In order to improve the agreement with the experimental data we have to include chiral oneโmeson loop corrections \[7โ9\] and the contribution of the $`\mathrm{\Delta }(1232)`$ resonance .
## 3 Chiral oneโmeson loop corrections to the amplitude of the M1โcapture
For the evaluation of chiral mesonโloop corrections in the NNJL we use (CHPT)<sub>q</sub> developed in Refs. within the ENJL model with a linear realization of chiral $`U(3)\times U(3)`$ symmetry. Below we consider the contributions of chiral oneโmeson loop corrections induced by the virtual meson transitions $`\pi a_1\gamma `$, $`a_1\pi \gamma `$, $`\pi (\omega ,\rho )\gamma `$, $`(\omega ,\rho )\pi \gamma `$, $`\sigma (\omega ,\rho )\gamma `$ and $`(\omega ,\rho )\sigma \gamma `$, where $`\sigma `$ is a scalar partner of pions under chiral $`SU(2)\times SU(2)`$ transformations in (CHPT)<sub>q</sub> with a linear realization of chiral $`U(3)\times U(3)`$ symmetry .
The effective Lagrangians $`\delta _{\mathrm{eff}}^{\mathrm{pp}\gamma }(x)`$ and $`\delta _{\mathrm{eff}}^{\mathrm{nn}\gamma }(x)`$, caused by the virtual meson transitions $`\pi a_1\gamma `$, $`a_1\pi \gamma `$, $`\pi (\omega ,\rho )\gamma `$, $`(\omega ,\rho )\pi \gamma `$, $`\sigma (\omega ,\rho )\gamma `$ and $`(\omega ,\rho )\sigma \gamma `$, we evaluate at leading order in the large $`N_C`$ expansion . The results of the evaluation contain divergent contributions. In order to remove these divergences we apply the renormalization procedure developed in (CHPT)<sub>q</sub> for the evaluation of chiral mesonโloop corrections (see Ivanov in Refs. ). Since the renormalized expressions should vanish in the chiral limit $`M_\pi 0`$ , only the virtual meson transitions with intermediate $`\pi `$โmeson give nonโtrivial contributions. The contributions of the virtual meson transitions with the intermediate $`\sigma `$โmeson are found finite in the chiral limit and subtracted according to the renormalization procedure . Such a cancellation of the $`\sigma `$โmeson contributions in the oneโmeson loop approximation agrees with Chiral perturbation theory using a nonโlinear realization of chiral symmetry, where $`\sigma `$โmesonโlike exchanges can appear only in twoโmeson loop approximation. Then, the sum of the contributions of the virtual meson transitions $`\pi ^{}\rho ^{}\gamma `$, $`\pi ^0\rho ^0\gamma `$ and $`\pi ^0\omega \gamma `$ to the effective coupling $`\mathrm{nn}\gamma `$ is equal to zero. As a result the effective Lagrangians $`\delta _{\mathrm{eff}}^{\mathrm{pp}\gamma }(x)`$ and $`\delta _{\mathrm{eff}}^{\mathrm{nn}\gamma }(x)`$ are given by
$`\delta _{\mathrm{eff}}^{\mathrm{pp}\gamma }(x)`$ $`=`$ $`{\displaystyle \frac{ie}{4M_\mathrm{N}}}\left[g_\mathrm{A}g_{\pi \mathrm{NN}}{\displaystyle \frac{\alpha _\rho }{16\pi ^3}}{\displaystyle \frac{M_\mathrm{N}}{F_\pi }}M_\pi ^2J_{\pi \mathrm{a}_1\mathrm{N}}+g_{\pi \mathrm{NN}}{\displaystyle \frac{N_C\alpha _\rho }{16\pi ^3}}{\displaystyle \frac{M_\mathrm{N}}{F_\pi }}M_\pi ^2J_{\pi \mathrm{VN}}\right],`$
$`\times [\overline{p}(x)\sigma _{\mu \nu }p(x)]F^{\mu \nu }(x),`$
$`\delta _{\mathrm{eff}}^{\mathrm{nn}\gamma }(x)`$ $`=`$ $`{\displaystyle \frac{ie}{4M_\mathrm{N}}}\left[g_\mathrm{A}g_{\pi \mathrm{NN}}{\displaystyle \frac{\alpha _\rho }{16\pi ^3}}{\displaystyle \frac{M_\mathrm{N}}{F_\pi }}M_\pi ^2J_{\pi \mathrm{a}_1\mathrm{N}}\right][\overline{n}(x)\sigma _{\mu \nu }n(x)]F^{\mu \nu }(x),`$ (3.1)
where $`\alpha _\rho =g_\rho ^2/4\pi =2.91`$ is the effective coupling constant of the $`\rho \pi \pi `$ decay, $`F_\pi =92.4\mathrm{MeV}`$ is the leptonic coupling constant of pions, and $`g_\mathrm{A}=1.267`$ . Then, $`J_{\pi \mathrm{a}_1\mathrm{N}}`$ and$`J_{\pi \mathrm{VN}}`$ are the momentum integrals determined by
$`J_{\pi \mathrm{a}_1\mathrm{N}}`$ $`=`$ $`{\displaystyle \frac{d^4p}{\pi ^2}\frac{1}{(M_\pi ^2+p^2)(M_{a_1}^2+p^2)(M_\mathrm{N}^2+p^2)}}=0.017M_\pi ^2,`$
$`J_{\pi \mathrm{VN}}`$ $`=`$ $`{\displaystyle \frac{d^4p}{\pi ^2}\frac{1}{(M_\pi ^2+p^2)(M_\mathrm{V}^2+p^2)(M_\mathrm{N}^2+p^2)}}=0.024M_\pi ^2,`$ (3.2)
where $`p`$ is Euclidean 4โmomentum, $`M_\mathrm{V}=M_\rho =M_\omega =770\mathrm{MeV}`$ and $`M_{a_1}=\sqrt{2}M_\rho `$ .
At $`N_C=3`$ the cross section for the M1โcapture accounting for the contribution of the effective interaction Eq.(3) amounts to
$`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})={\displaystyle \frac{1}{v}}(\mu _\mathrm{p}\mu _\mathrm{n})^2{\displaystyle \frac{25}{64}}{\displaystyle \frac{\alpha }{\pi ^2}}Q_\mathrm{D}C_{\mathrm{NN}}^2M_\mathrm{N}\epsilon _\mathrm{D}^3`$
$`\times [1+{\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{M_\pi ^2}{8\pi ^2}}{\displaystyle \frac{\alpha _\rho }{\pi }}(J_{\pi \mathrm{a}_1\mathrm{N}}+{\displaystyle \frac{3}{2g_\mathrm{A}}}J_{\pi \mathrm{VN}})]^2=287.2\mathrm{mb},`$ (3.3)
where we have used the relation $`g_{\pi \mathrm{NN}}g_\mathrm{A}M_\mathrm{N}/F_\pi `$. The theoretical value of the cross section for the neutronโproton radiative capture given by Eq.(3) differs from the experimental one by about 14$`\%`$. This discrepancy we should describe by taking into account the contribution of the $`\mathrm{\Delta }(1232)`$ resonance.
## 4 The $`\mathrm{\Delta }(1232)`$ resonance
For the evaluation of the contribution of the $`\mathrm{\Delta }(1232)`$ resonance to the amplitude of the M1โcapture in the NNJL model we have to obtain the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}}(x)`$ describing the strong lowโenergy n + p $``$ $`\mathrm{\Delta }`$ \+ N transition. For this aim we should use the procedure having been applied to the evaluation of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{np}}(x)`$ given by Eq.(1). In Refs. the evaluation of the contribution of the $`\mathrm{\Delta }(1232)`$ resonance in terms of exchange currents has been carried out in the oneโpion exchange approximation. Thereby, following Refs. we suppose to evaluate the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}}(x)`$ of the strong lowโenergy n + p $``$ $`\mathrm{\Delta }`$ \+ N transition in the oneโpion exchange approximation. This gives
$`_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}}(x)={\displaystyle \frac{i}{\sqrt{6}}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{M_\mathrm{N}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}}{4M_\pi ^2}}{\displaystyle }d^4z{\displaystyle \frac{}{z_\phi }}\delta ^{(4)}(zx)\{[\overline{\mathrm{\Delta }}_\omega ^+(z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}n^c(x)]`$
$`\times [\overline{n^c}(z)\gamma ^5p(x)+\overline{n^c}(x)\gamma ^5p(z)][\overline{\mathrm{\Delta }}_\omega ^0(z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}p^c(x)][\overline{n^c}(z)\gamma ^5p(x)+\overline{n^c}(x)\gamma ^5p(z)]`$
$`+1\gamma ^5\gamma _\nu \gamma ^\nu \gamma ^5\},`$ (4.1)
where we have kept only the terms contributing to the transition of the np pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state into the $`\mathrm{\Delta }N`$ state. Using then the phenomenological Lagrangian
$`_{\mathrm{npD}}(x)=ig_\mathrm{V}[\overline{p^c}(x)\gamma ^\mu n(x)\overline{n^c}(x)\gamma ^\mu p(x)]D_\mu ^{}(x)`$ (4.2)
the effective Lagrangian describing the contribution of the $`\mathrm{\Delta }(1232)`$ resonance to the lowโenergy transition n + p $``$ D + $`\gamma `$ is defined by
$`{\displaystyle d^4x_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}\mathrm{D}\gamma }(x)}={\displaystyle d^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}}(x_1)_{\mathrm{npD}}(x_2)_{\gamma \mathrm{N}\mathrm{\Delta }}(x_3))}=`$
$`={\displaystyle \frac{i}{6}}{\displaystyle \frac{eg_\mathrm{V}}{M_\mathrm{N}^2}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}{\displaystyle d^4x_1d^4x_2d^4x_3d^4z\frac{}{z_\phi }\delta ^{(4)}(zx_1)}`$
$`\times \mathrm{T}([\overline{p^c}(x_1)\gamma ^5n(z)+\overline{p^c}(z)\gamma ^5n(x_1)]D_\mu ^{}(x_2)F^{\alpha \beta }(x_3))`$
$`\times \{0|\mathrm{T}([\overline{\mathrm{\Delta }}_\omega ^+(z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}n^c(x_1)][\overline{p^c}(x_2)\gamma ^\mu n(x_2)\overline{n^c}(x_2)\gamma ^\mu p(x_2)][\overline{p}(x_3)\gamma _\beta \gamma ^5\mathrm{\Delta }_\alpha ^+(x_3)])|0`$
$`0|\mathrm{T}([\overline{\mathrm{\Delta }}_\omega ^0(z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}p^c(x)][\overline{p^c}(x_2)\gamma ^\mu n(x_2)\overline{n^c}(x_2)\gamma ^\mu p(x_2)]`$
$`\times [\overline{n}(x_3)\gamma _\beta \gamma ^5\mathrm{\Delta }_\alpha ^0(x_3)])|0+(\gamma ^51\gamma _\nu \gamma ^5\gamma ^\nu )\}=`$
$`={\displaystyle \frac{i}{3}}{\displaystyle \frac{eg_\mathrm{V}}{M_\mathrm{N}}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}{\displaystyle d^4x_1d^4x_2d^4x_3d^4z\frac{}{z_\phi }\delta ^{(4)}(zx_1)}`$
$`\times \mathrm{T}([\overline{p^c}(x_1)\gamma ^5n(z)+\overline{p^c}(z)\gamma ^5n(x_1)]D_\mu ^{}(x_2)F^{\alpha \beta }(x_3))`$
$`\times \{0|\mathrm{T}([\overline{\mathrm{\Delta }}_\omega ^+(z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}n^c(x_1)][\overline{n^c}(x_2)\gamma ^\mu p(x_2)][\overline{p}(x_3)\gamma _\beta \gamma ^5\mathrm{\Delta }_\alpha ^+(x_3)])|0`$
$`+0|\mathrm{T}([\overline{\mathrm{\Delta }}_\omega ^0(z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}p^c(x_1)][\overline{p^c}(x_2)\gamma ^\mu n(x_2)][\overline{n}(x_3)\gamma _\beta \gamma ^5\mathrm{\Delta }_\alpha ^0(x_3)])|0`$
$`+(\gamma ^51\gamma _\nu \gamma ^5\gamma ^\nu )\}=`$
$`={\displaystyle \frac{2}{3}}{\displaystyle \frac{ieg_\mathrm{V}}{M_\mathrm{N}^2}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}{\displaystyle d^4x_1d^4x_2d^4x_3d^4z\frac{}{z_\phi }\delta ^{(4)}(zx_1)}`$
$`\times \{\mathrm{T}([\overline{p^c}(x_1)\gamma ^5n(z)+\overline{p^c}(z)\gamma ^5n(x_1)]D_\mu ^{}(x_2)F^{\alpha \beta }(x_3))`$
$`\times {\displaystyle \frac{1}{i}}\mathrm{tr}\{S_{\alpha \omega }(x_3z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}S_\mathrm{F}^c(x_1x_2)\gamma ^\mu S_\mathrm{F}(x_2x_3)\gamma _\beta \gamma ^5\}`$
$`\mathrm{T}([\overline{p^c}(x_1)\gamma _\nu \gamma ^5n(z)+\overline{p^c}(z)\gamma _\nu \gamma ^5n(x_1)]D_\mu ^{}(x_2)F^{\alpha \beta }(x_3))`$
$`\times {\displaystyle \frac{1}{i}}\mathrm{tr}\{S_{\alpha \omega }(x_3z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}\gamma ^\nu S_\mathrm{F}^c(x_1x_2)\gamma ^\mu S_\mathrm{F}(x_2x_3)\gamma _\beta \gamma ^5\}\}.`$ (4.3)
Thus, the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}\mathrm{D}\gamma }(x)`$ is equal to
$`{\displaystyle d^4x_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}\mathrm{D}\gamma }(x)}={\displaystyle \frac{2}{3}}{\displaystyle \frac{ieg_\mathrm{V}}{M_\mathrm{N}^2}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}{\displaystyle d^4x_1d^4x_2d^4x_3d^4z\frac{}{z_\phi }\delta ^{(4)}(zx_1)}`$
$`\times \{\mathrm{T}([\overline{p^c}(x_1)\gamma ^5n(z)+\overline{p^c}(z)\gamma ^5n(x_1)]D_\mu ^{}(x_2)F^{\alpha \beta }(x_3))`$
$`\times {\displaystyle \frac{1}{i}}\mathrm{tr}\{S_{\alpha \omega }(x_3z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}S_\mathrm{F}^c(x_1x_2)\gamma ^\mu S_\mathrm{F}(x_2x_3)\gamma _\beta \gamma ^5\}`$
$`\mathrm{T}([\overline{p^c}(x_1)\gamma _\nu \gamma ^5n(z)+\overline{p^c}(z)\gamma _\nu \gamma ^5n(x_1)]D_\mu ^{}(x_2)F^{\alpha \beta }(x_3))`$
$`\times {\displaystyle \frac{1}{i}}\mathrm{tr}\{S_{\alpha \omega }(x_3z)\mathrm{\Theta }_{}^{\omega }{}_{\phi }{}^{}\gamma ^\nu S_\mathrm{F}^c(x_1x_2)\gamma ^\mu S_\mathrm{F}(x_2x_3)\gamma _\beta \gamma ^5\}\}.`$ (4.4)
In the momentum representation of the baryon Green functions the effective Lagrangian Eq.(4) reads
$`{\displaystyle d^4x_{\mathrm{eff}}^{\mathrm{np}\mathrm{\Delta }\mathrm{N}\mathrm{D}\gamma }(x)}={\displaystyle \frac{2}{3}}{\displaystyle \frac{ieg_\mathrm{V}}{M_\mathrm{N}^2}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}{\displaystyle d^4x_1d^4z\frac{}{z_\phi }\delta ^{(4)}(zx_1)}`$
$`\times {\displaystyle }{\displaystyle \frac{d^4x_2d^4k_2}{(2\pi )^4}}{\displaystyle \frac{d^4x_3d^4k_3}{(2\pi )^4}}e^{ik_2\left(x_2x_1\right)}e^{ik_3\left(x_3z\right)}`$
$`\times \{\mathrm{T}([\overline{p^c}(x_1)\gamma ^5n(z)+\overline{p^c}(z)\gamma ^5n(x_1)]D_\mu ^{}(x_2)F_{\alpha \beta }(x_3))`$
$`\times {\displaystyle }{\displaystyle \frac{d^4k_1}{\pi ^2i}}e^{ik_1\left(x_1z\right)}\mathrm{tr}\{S^{\alpha \omega }(k_1+k_3)\mathrm{\Theta }_{\omega \phi }{\displaystyle \frac{1}{M_\mathrm{N}\widehat{k}_1+\widehat{k}_2}}\gamma ^\mu {\displaystyle \frac{1}{M_\mathrm{N}\widehat{k}_1}}\gamma ^\beta \gamma ^5\}`$
$`\mathrm{T}([\overline{p^c}(x_1)\gamma _\nu \gamma ^5n(z)+\overline{p^c}(z)\gamma _\nu \gamma ^5n(x_1)]D_\mu ^{}(x_2)F^{\alpha \beta }(x_3))`$
$`\times {\displaystyle }{\displaystyle \frac{d^4k_1}{\pi ^2i}}e^{ik_1\left(x_1z\right)}\mathrm{tr}\{S^{\alpha \omega }(k_1+k_3)\mathrm{\Theta }_{\omega \phi }{\displaystyle \frac{1}{M_\mathrm{N}\widehat{k}_1+\widehat{k}_2}}\gamma ^\mu {\displaystyle \frac{1}{M_\mathrm{N}\widehat{k}_1}}\gamma ^\beta \gamma ^5\}\}.`$ (4.5)
The effective Lagrangian Eq.(4) defines the contribution of the $`\mathrm{\Delta }(1232)`$ resonance to the lowโenergy transition n + p $``$ D + $`\gamma `$.
The matrix element of the neutronโproton radiative capture caused by the contribution of the $`\mathrm{\Delta }(1232)`$ resonance exchange is equal to
$`(\mathrm{n}+\mathrm{p}\mathrm{\Delta }\mathrm{N}\mathrm{D}+\gamma )=`$
$`={\displaystyle \frac{ie}{2M_\mathrm{N}^2}}{\displaystyle \frac{g_\mathrm{V}}{6\pi ^2}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}[\overline{u^c}(p_2)\gamma ^5u(p_1)](k_\alpha e_\beta ^{}(k)k_\beta e_\alpha ^{}(k))e_\mu ^{}(k_\mathrm{D})`$
$`\times ๐ฅ_5^{\mu \beta \alpha }(k_\mathrm{D},k)`$
$`+{\displaystyle \frac{ie}{2M_\mathrm{N}^2}}{\displaystyle \frac{g_\mathrm{V}}{6\pi ^2}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}[\overline{u^c}(p_2)\gamma _\nu \gamma ^5u(p_1)](k_\alpha e_\beta ^{}(k)k_\beta e_\alpha ^{}(k))e_\mu ^{}(k_\mathrm{D})`$
$`\times ๐ฅ_5^{\nu \mu \beta \alpha }(k_\mathrm{D},k),`$ (4.6)
where the structure functions $`๐ฅ_5^{\mu \beta \alpha }(k_\mathrm{D},k)`$ and $`๐ฅ_5^{\nu \mu \beta \alpha }(k_\mathrm{D},k)`$ are defined by the momentum integrals
$`๐ฅ_5^{\mu \beta \alpha }(k_\mathrm{D},k)=`$
$`={\displaystyle \frac{d^4k_1}{\pi ^2i}\mathrm{tr}\{(k_1+k_\mathrm{D})^\phi S^{\alpha \omega }(k_1+k)\mathrm{\Theta }_{\omega \phi }\frac{1}{M_\mathrm{N}\widehat{k}_1+\widehat{k}_\mathrm{D}}\gamma ^\mu \frac{1}{M_\mathrm{N}\widehat{k}_1}\gamma ^\beta \gamma ^5\}},`$
$`๐ฅ_5^{\nu \mu \beta \alpha }(k_\mathrm{D},k)=`$
$`={\displaystyle \frac{d^4k_1}{\pi ^2i}\mathrm{tr}\{(k_1+k_\mathrm{D})^\phi S^{\alpha \omega }(k_1+k)\mathrm{\Theta }_{\omega \phi }\gamma ^\nu \frac{1}{M_\mathrm{N}\widehat{k}_1+\widehat{k}_\mathrm{D}}\gamma ^\mu \frac{1}{M_\mathrm{N}\widehat{k}_1}\gamma ^\beta \gamma ^5\}}.`$ (4.7)
At leading order in the large $`N_C`$ expansion the structure functions Eq.(4) read
$`๐ฅ_5^{\mu \beta \alpha }(k_\mathrm{D},k)`$ $`=`$ $`{\displaystyle \frac{4}{3}}\left(Z{\displaystyle \frac{1}{2}}\right)iM_\mathrm{N}\epsilon ^{\mu \beta \alpha \lambda }k_{\mathrm{D}\lambda },`$
$`๐ฅ_5^{\nu \mu \beta \alpha }(k_\mathrm{D},k)`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(Z{\displaystyle \frac{1}{2}}\right)iM_\mathrm{N}^2\epsilon ^{\mu \beta \alpha \nu }.`$ (4.8)
We have neglected the mass difference between the masses of the $`\mathrm{\Delta }(1232)`$ resonance and the nucleon. The matrix element of the lowโenergy transition n + p $``$ D + $`\gamma `$ caused by the $`\mathrm{\Delta }(1232)`$ resonance contribution is equal to
$`_\mathrm{\Delta }(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )={\displaystyle \frac{e}{2M_\mathrm{N}}}{\displaystyle \frac{g_\mathrm{V}}{4\pi ^2}}\left[\left({\displaystyle \frac{1}{2}}Z\right){\displaystyle \frac{8}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}\right]`$
$`\times \epsilon ^{\alpha \beta \mu \nu }k_\alpha e_\beta ^{}(k,\lambda )e_\mu ^{}(k_\mathrm{D},\lambda _\mathrm{D})[\overline{u^c}(p_2)(2k_{\mathrm{D}\nu }M_\mathrm{N}\gamma _\nu )\gamma ^5u(p_1)]=`$
$`=e{\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}\left[\left({\displaystyle \frac{1}{2}}Z\right){\displaystyle \frac{8}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}\right](\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})`$
$`\times [\overline{u^c}(p_2)\gamma ^5u(p_1)].`$ (4.9)
In turn, the contribution of the nucleon tensor current
$`\delta _{\mathrm{npD}}(x)={\displaystyle \frac{g_\mathrm{T}}{2M_\mathrm{N}}}[\overline{p^c}(x)\sigma ^{\mu \nu }n(x)\overline{n^c}(x)\sigma ^{\mu \nu }p(x)]D_{\mu \nu }^{}(x)`$ (4.10)
does not depend on the parameter $`Z`$ and reads
$`\delta _\mathrm{\Delta }(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )=e{\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}\left[{\displaystyle \frac{1}{5}}{\displaystyle \frac{g_\mathrm{T}}{g_\mathrm{V}}}{\displaystyle \frac{8}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}\right]`$ (4.11)
$`\times (\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})[\overline{u^c}(p_2)\gamma ^5u(p_1)].`$
The coupling constants $`g_\mathrm{T}`$ and $`g_\mathrm{V}`$ are connected by the relation
$`g_\mathrm{T}=\sqrt{{\displaystyle \frac{3}{8}}}g_\mathrm{V}+O(1/\sqrt{N_C}).`$ (4.12)
The total amplitude of the neutronโproton radiative capture for thermal neutrons reads
$`(\mathrm{n}+\mathrm{p}\mathrm{D}+\gamma )=e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}^{}(\stackrel{}{k},\lambda ))\stackrel{}{e}^{}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})[\overline{u^c}(p_2)\gamma ^5u(p_1)]`$
$`\times [1+{\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{M_\pi ^2}{8\pi ^2}}{\displaystyle \frac{\alpha _\rho }{\pi }}(J_{\pi \mathrm{a}_1\mathrm{N}}+{\displaystyle \frac{3}{2g_\mathrm{A}}}J_{\pi \mathrm{VN}})+{\displaystyle \frac{1}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{5}}\sqrt{{\displaystyle \frac{3}{8}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{8}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}`$
$`+{\displaystyle \frac{12Z}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{4}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}].`$ (4.13)
The total cross section for the neutronโproton radiative capture is then defined by
$`\sigma (\mathrm{np}\mathrm{D}\gamma )(p)={\displaystyle \frac{1}{v}}(\mu _\mathrm{p}\mu _\mathrm{n})^2{\displaystyle \frac{25}{64}}{\displaystyle \frac{\alpha }{\pi ^2}}Q_\mathrm{D}C_{\mathrm{NN}}^2M_\mathrm{N}\epsilon _\mathrm{D}^3`$
$`\times [1+{\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{M_\pi ^2}{8\pi ^2}}{\displaystyle \frac{\alpha _\rho }{\pi }}(J_{\pi \mathrm{a}_1\mathrm{N}}+{\displaystyle \frac{3}{2g_\mathrm{A}}}J_{\pi \mathrm{VN}})+{\displaystyle \frac{1}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{5}}\sqrt{{\displaystyle \frac{3}{8}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{8}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}`$
$`+{\displaystyle \frac{12Z}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{4}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}]^2.`$ (4.14)
The numerical value of the cross section amounts to
$`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=325.5(1+0.246(12Z))^2\mathrm{mb}.`$ (4.15)
Thus, the discrepancy of the theoretical cross section and the experimental value Eq.(1.1) can by described by the contribution of the $`\mathrm{\Delta }(1232)`$ resonance. In order to fit the experimental value of the cross section we should set $`Z=0.473`$. This agrees with the experimental bound $`|Z|1/2`$ . At $`Z=1/2`$ that is favoured theoretically we get the cross section $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=325.5\mathrm{mb}`$ agreeing with the experimental value with accuracy better than 3$`\%`$.
## 5 The photomagnetic disintegration of the deuteron
The amplitude of the photomagnetic disintegration of the deuteron $`\gamma `$ \+ D $``$ n + p is related to the amplitude of the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ due to timeโreversal invariance and reads
$`(\gamma +\mathrm{D}\mathrm{n}+\mathrm{p})=e(\mu _\mathrm{p}\mu _\mathrm{n}){\displaystyle \frac{5g_\mathrm{V}}{8\pi ^2}}C_{\mathrm{NN}}(\stackrel{}{k}\times \stackrel{}{e}(\stackrel{}{k},\lambda ))\stackrel{}{e}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})[\overline{u}(p_2)\gamma ^5u^c(p_1)]`$
$`\times [1+{\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{M_\pi ^2}{8\pi ^2}}{\displaystyle \frac{\alpha _\rho }{\pi }}(J_{\pi \mathrm{a}_1\mathrm{N}}+{\displaystyle \frac{3}{2g_\mathrm{A}}}J_{\pi \mathrm{VN}})+{\displaystyle \frac{1}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{5}}\sqrt{{\displaystyle \frac{3}{8}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{8}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}`$
$`+{\displaystyle \frac{12Z}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{4}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}]^2e^{i\delta _{\mathrm{np}}\left(p\right)}{\displaystyle \frac{\mathrm{sin}\delta _{\mathrm{np}}(p)}{a_{\mathrm{np}}p}}{\displaystyle \frac{1}{1+r_\mathrm{D}^2p^2}}.`$ (5.1)
The cross section defined by the amplitude Eq.(5) is then given by
$`\sigma (\gamma \mathrm{D}\mathrm{np})(\omega )=\sigma _0\left({\displaystyle \frac{\omega }{\epsilon _\mathrm{D}}}\right){\displaystyle \frac{1}{\left(1{\displaystyle \frac{1}{2}}r_{\mathrm{np}}a_{\mathrm{np}}p^2\right)^2+a_{\mathrm{np}}^2p^2}}{\displaystyle \frac{r_\mathrm{D}p}{(1+r_\mathrm{D}^2p^2)^2}},`$ (5.2)
where $`p=\sqrt{M_\mathrm{N}(\omega \epsilon _\mathrm{D})}`$ is the relative momentum of the np pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state, $`\omega `$ is the energy of the photon, and $`\sigma _0`$ is equal to
$`\sigma _0=(\mu _\mathrm{p}\mu _\mathrm{n})^2{\displaystyle \frac{25\alpha Q_\mathrm{D}}{192\pi ^2}}C_{\mathrm{NN}}^2\epsilon _\mathrm{D}^{3/2}M_\mathrm{N}^{5/2}`$
$`\times [1+{\displaystyle \frac{g_{\pi \mathrm{NN}}^2}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{M_\pi ^2}{8\pi ^2}}{\displaystyle \frac{\alpha _\rho }{\pi }}(J_{\pi \mathrm{a}_1\mathrm{N}}+{\displaystyle \frac{3}{2g_\mathrm{A}}}J_{\pi \mathrm{VN}})+{\displaystyle \frac{1}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{5}}\sqrt{{\displaystyle \frac{3}{8}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{8}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}`$
$`+{\displaystyle \frac{12Z}{\mu _\mathrm{p}\mu _\mathrm{n}}}{\displaystyle \frac{1}{C_{\mathrm{NN}}}}{\displaystyle \frac{4}{9}}{\displaystyle \frac{g_{\pi \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\gamma \mathrm{N}\mathrm{\Delta }}}{g_{\pi \mathrm{NN}}}}{\displaystyle \frac{g_{\pi \mathrm{NN}}^3}{4M_\pi ^2}}]^2=7.10\mathrm{mb}.`$ (5.3)
The energy region of the dominance of the photomagnetic disintegration of the deuteron is restricted by the constraint $`r_\mathrm{D}p=\sqrt{(\omega \epsilon _\mathrm{D})/\epsilon _\mathrm{D}}1`$ or differently $`\omega \epsilon _\mathrm{D}=2.225\mathrm{MeV}`$.
The numerical values of the cross section for the photomagnetic disintegration of the deuteron at energies $`\omega 2\epsilon _\mathrm{D}=4.45\mathrm{MeV}`$ read
$`\sigma (\gamma \mathrm{D}\mathrm{np})(\omega )|_{\omega =2.62\mathrm{MeV}}`$ $`=`$ $`0.358(0.380)\mathrm{mb},`$
$`\sigma (\gamma \mathrm{D}\mathrm{np})(\omega )|_{\omega =2.76\mathrm{MeV}}`$ $`=`$ $`0.302(0.327)\mathrm{mb},`$
$`\sigma (\gamma \mathrm{D}\mathrm{np})(\omega )|_{\omega =4.45\mathrm{MeV}}`$ $`=`$ $`0.094(0.128)\mathrm{mb},`$ (5.4)
where in parentheses we have adduced the theoretical values obtained by Chen and Savage in the EFT . One can see a reasonable agreement between the results obtained in the NNJL model and the EFT. Thus, the spatial smearing of the physical deuteron caused by the effective radius $`r_\mathrm{D}`$ and introduced in the NNJL model phenomenologically in the form of the wave function Eq.(2.16) describes well the energy dependence of the cross section for the photomagnetic disintegration of the deuteron at photon energies far from threshold. Note that in the critical region of the photon energies $`\omega 2\epsilon _\mathrm{D}=4.45\mathrm{MeV}`$ the cross section for the photomagnetic disintegration of the deuteron calculated in the NNJL model falls steeper with $`\omega `$ than in the EFT. However, in this energy region the dominant role is attributed to the E1โtransition which we have not taken into account.
For the correct description of the experimental data on the photodisintegration of the deuteron (see also Chen and Savage ):
$`\sigma (\gamma \mathrm{D}\mathrm{np})_{\mathrm{exp}}(\omega )|_{\omega =2.62\mathrm{MeV}}`$ $`=`$ $`(1.300\pm 0.029)\mathrm{mb},`$
$`\sigma (\gamma \mathrm{D}\mathrm{np})_{\mathrm{exp}}(\omega )|_{\omega =2.76\mathrm{MeV}}`$ $`=`$ $`(1.474\pm 0.032)\mathrm{mb},`$
$`\sigma (\gamma \mathrm{D}\mathrm{np})_{\mathrm{exp}}(\omega )|_{\omega =4.45\mathrm{MeV}}`$ $`=`$ $`(2.430\pm 0.170)\mathrm{mb}`$ (5.5)
one must include the contribution of the E1โtransition .
Since the cross section for the photodisintegration of the deuteron having been discussed above does not contain the contribution of the E1โtransition, our result can be scarcely compared with the recent theoretical investigation of the photodisintegration of the deuteron carried out by Anisovich and Sadovnikova within the dispersion relation approach based on the dispersion relation technique developed by Anisovich et al. . This investigation represents a detailed analysis of the saturation of the cross section for the photodisintegration of the deuteron by different intermediate states valid mainly for the photon energy region $`\omega 50\mathrm{MeV}`$, where the contribution of the E1โtransition is important.
## 6 The amplitude of the solar proton burning
The analysis of the reaction p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ within the NNJL model we should start with the evaluation of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ of the transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$. In the treeโmeson approximation<sup>1</sup><sup>1</sup>1Below the analysis of weak nuclear reactions is carried out in the treeโmeson approximation. The inclusion of chiral oneโmeson loop corrections goes beyond the scope of this paper and demands a separate publication. the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ is defined by the oneโnucleon loop diagrams depicted in Fig.3. The detailed evaluation of $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ is given in Appendix. The result reads<sup>2</sup><sup>2</sup>2This result is obtained at zero contribution of the nucleon tensor current (see Appendix and discussion in the Conclusion).
$`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=ig_\mathrm{A}C_{\mathrm{NN}}M_\mathrm{N}{\displaystyle \frac{G_\mathrm{V}}{\sqrt{2}}}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}D_\mu ^{}(x)[\overline{p^c}(x)\gamma ^5p(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\mu (1\gamma ^5)\psi _\mathrm{e}(x)].`$ (6.1)
where $`G_\mathrm{V}=G_\mathrm{F}\mathrm{cos}\vartheta _C`$ with $`G_\mathrm{F}=1.166\times \mathrm{\hspace{0.17em}10}^{11}\mathrm{MeV}^2`$ and $`\vartheta _C`$ are the Fermi weak coupling constant and the Cabibbo angle $`\mathrm{cos}\vartheta _C=0.975`$
For the derivation of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ we have used the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{NN}\mathrm{NN}}(x)`$ responsible for lowโenergy transitions N + N $``$ N + N defined by Eq.(1).
The matrix element of the transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ we define by a usual way
$`{\displaystyle d^4xD(k_\mathrm{D})e^+(k_{\mathrm{e}^+})\nu _\mathrm{e}(k_{\nu _\mathrm{e}})|_{\mathrm{eff}}^{\mathrm{np}\mathrm{De}^+\nu _\mathrm{e}}(x)|p(p_1)p(p_2)}=`$
$`=(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\mathrm{e}^+}+k_{\nu _\mathrm{e}}p_1p_2){\displaystyle \frac{(\mathrm{n}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _\mathrm{e})}{\sqrt{2E_1V\mathrm{\hspace{0.17em}2}E_2V\mathrm{\hspace{0.17em}2}E_\mathrm{D}V\mathrm{\hspace{0.17em}2}E_{\mathrm{e}^+}V\mathrm{\hspace{0.17em}2}E_{\nu _\mathrm{e}}V}}},`$ (6.2)
where $`E_i(i=1,2,\mathrm{D},\mathrm{e}^+,\nu _\mathrm{e})`$ are the energies of the protons, the deuteron, the positron and the neutrino, $`V`$ is the normalization volume.
The wave functions of the initial $`|p(p_1)p(p_2)`$ and final $`D(k_\mathrm{D})e^+(k_{\mathrm{e}^+})\nu _\mathrm{e}(k_{\nu _\mathrm{e}})|`$ state we take in the usual form
$`|p(p_1)p(p_2)`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}a_\mathrm{p}^{}(\stackrel{}{p}_1,\sigma _1)a_\mathrm{p}^{}(\stackrel{}{p}_2,\sigma _2)|0,`$
$`D(k_\mathrm{D})e^+(k_{\mathrm{e}^+})\nu _\mathrm{e}(k_{\nu _\mathrm{e}})|`$ $`=`$ $`0|a_\mathrm{D}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})a_{\mathrm{e}^+}(\stackrel{}{k}_{\mathrm{e}^+},\sigma )a_{\nu _\mathrm{e}}(\stackrel{}{k}_{\nu _\mathrm{e}},\sigma ^{}),`$ (6.3)
where $`a_\mathrm{n}^{}(\stackrel{}{p}_1,\sigma _1)`$ and $`a_\mathrm{p}^{}(\stackrel{}{p}_2,\sigma _2)`$ are the operators of creation of the protons. In turn, $`a_\mathrm{D}(\stackrel{}{k}_\mathrm{D},\lambda _\mathrm{D})`$, $`a_{\mathrm{e}^+}(\stackrel{}{k}_{\mathrm{e}^+},\sigma _{\mathrm{e}^+})`$ and $`a_{\nu _\mathrm{e}}(\stackrel{}{k}_{\nu _\mathrm{e}},\sigma _{\nu _\mathrm{e}})`$ are the operators of annihilation of the deuteron, the positron and the neutrino.
The effective Lagrangian Eq.(6.1) defines the effective vertex of the lowโenergy nuclear transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$
$`i(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)`$ $`=`$ $`g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}G_\mathrm{V}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}})\gamma ^\mu (1\gamma ^5)v(k_{\mathrm{e}^+})]`$ (6.4)
$`\times [\overline{u^c}(p_2)\gamma ^5u(p_1)],`$
where $`e_\mu ^{}(k_\mathrm{D},\lambda _\mathrm{D})`$ is a 4โvector of a polarization of the deuteron, $`u(k_{\nu _\mathrm{e}})`$, $`v(k_{\mathrm{e}^+})`$, $`u(p_2)`$ and $`u(p_1)`$ are the Dirac bispinors of the neutrino, the positron and the protons, respectively.
For the evaluation of the matrix element Eq.(6.4) we have used the wave functions of the protons in the form of the plane waves. However, a real wave function of the pp pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state is defined by lowโenergy nuclear forces and Coloumb repulsion. In order to take into account both lowโenergy nuclear forces and Coulomb repulsion for the relative movement of the pp pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state we would sum up an infinite series of oneโproton bubbles. The vertices of these oneโnucleon bubbles are defined by
$`V_{\mathrm{pp}\mathrm{pp}}(k^{},k)=C_{\mathrm{NN}}\psi _{\mathrm{pp}}^{}(k^{})[\overline{u}(p_2^{})\gamma ^5u^c(p_1^{})][\overline{u^c}(p_2)\gamma ^5u(p_1)]\psi _{\mathrm{pp}}(k),`$ (6.5)
where $`\psi _{\mathrm{pp}}(k)`$ and $`\psi _{\mathrm{pp}}^{}(k^{})`$ are the explicit Coulomb wave functions of the relative movement of the protons taken at zero relative distances, and $`k`$ and $`k^{}`$ are relative 3โmomenta of the protons $`\stackrel{}{k}=(\stackrel{}{p}_1\stackrel{}{p}_2)/2`$ and $`\stackrel{}{k}^{}=(\stackrel{}{p}_1^{}\stackrel{}{p}_2^{})/2`$ in the initial and final states. The explicit form of $`\psi _{\mathrm{pp}}(k)`$ we take following Kong and Ravndal (see also )
$`\psi _{\mathrm{pp}}(k)=e^{\pi /4kr_C}\mathrm{\Gamma }\left(1+{\displaystyle \frac{i}{2kr_C}}\right),`$ (6.6)
where $`2r_C=2/M_\mathrm{N}\alpha =57.64\mathrm{fm}`$ and $`\alpha =1/137`$ are the Bohr radius of the pp pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state and the fine structure constant. The squared value of the modulo of $`\psi _{\mathrm{pp}}(k)`$ is given by
$`|\psi _{\mathrm{pp}}(k)|^2=C_0^2(k)={\displaystyle \frac{\pi }{kr_C}}{\displaystyle \frac{1}{e^{\pi /kr_C}1}},`$ (6.7)
where $`C_0(k)`$ is the Gamow penetration factor .
By taking into account the contribution of the Coulomb wave function and summing up an infinite series of oneโproton bubbles the expression Eq.(6.4) can be recast into the form
$`i(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)=g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}G_\mathrm{V}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}})\gamma ^\mu (1\gamma ^5)v(k_{\mathrm{e}^+})]`$
$`\times {\displaystyle \frac{[\overline{u^c}(p_2)\gamma ^5u(p_1)]\psi _{\mathrm{pp}}(k)}{1+{\displaystyle \frac{C_{\mathrm{NN}}}{16\pi ^2}}{\displaystyle \frac{d^4p}{\pi ^2i}|\psi _{\mathrm{pp}}(|\stackrel{}{p}+\stackrel{}{Q}|)|^2\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{p}\widehat{P}\widehat{Q}}\gamma ^5\frac{1}{M_\mathrm{N}\widehat{p}\widehat{Q}}\right\}}}}.`$ (6.8)
where $`P=p_1+p_2=(2\sqrt{k^2+M_\mathrm{N}^2},\stackrel{}{0})`$ is the 4โmomentum of the ppโpair in the center of mass frame; $`Q=aP+bK=a(p_1+p_2)+b(p_1p_2)`$ is an arbitrary shift of virtual momentum with arbitrary parameters $`a`$ and $`b`$, and in the center of mass frame $`K=p_1p_2=(0,2\stackrel{}{k})`$. The parameters $`a`$ and $`b`$ can be functions of $`k`$.
The evaluation of the momentum integral we would carry out at leading order in the $`1/M_\mathrm{N}`$ expansion or differently in the large $`N_C`$ expansion due to proportionality $`M_\mathrm{N}N_C`$ valid in QCD with $`SU(N_C)`$ gauge group at $`N_C\mathrm{}`$ . As a result we obtain
$`{\displaystyle \frac{d^4p}{\pi ^2i}|\psi _{\mathrm{pp}}(|\stackrel{}{p}+\stackrel{}{Q}|)|^2\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{p}\widehat{P}\widehat{Q}}\gamma ^5\frac{1}{M_\mathrm{N}\widehat{p}\widehat{Q}}\right\}}=`$
$`=8a(a+1)M_\mathrm{N}^2+8(b^2a(a+1))k^2i\mathrm{\hspace{0.17em}8}\pi M_\mathrm{N}k|\psi _{\mathrm{pp}}(k)|^2=`$
$`=8a(a+1)M_\mathrm{N}^2+8(b^2a(a+1))k^2i\mathrm{\hspace{0.17em}8}\pi M_\mathrm{N}kC_0^2(k).`$ (6.9)
Substituting Eq.(6) in Eq.(6) we get
$`i(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)=g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}G_\mathrm{V}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}`$
$`\times e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}})\gamma ^\mu (1\gamma ^5)v(k_{\mathrm{e}^+})][\overline{u^c}(p_2)\gamma ^5u(p_1)]e^{\pi /4kr_C}\mathrm{\Gamma }\left(1+{\displaystyle \frac{i}{2kr_C}}\right)`$
$`\left[1a(a+1){\displaystyle \frac{G_{\pi \mathrm{NN}}}{2\pi ^2}}M_\mathrm{N}^2+{\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}(b^2a(a+1))k^2i{\displaystyle \frac{C_{\mathrm{NN}}M_\mathrm{N}}{2\pi }}kC_0^2(k)\right]^1.`$ (6.10)
In order to reconcile the contribution of lowโenergy elastic pp scattering with lowโenergy nuclear phenomenology we should make a few changes. For this aim we should rewrite Eq.(6) in more convenient form
$`i(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)=g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}G_\mathrm{V}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}`$
$`\times e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}})\gamma ^\mu (1\gamma ^5)v(k_{\mathrm{e}^+})][\overline{u^c}(p_2)\gamma ^5u(p_1)]e^{i\sigma _0\left(k\right)}C_0(k)`$
$`\left[1a(a+1){\displaystyle \frac{G_{\pi \mathrm{NN}}}{2\pi ^2}}M_\mathrm{N}^2+{\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}(b^2a(a+1))k^2i{\displaystyle \frac{C_{\mathrm{NN}}M_\mathrm{N}}{2\pi }}kC_0^2(k)\right]^1.`$ (6.11)
We have denoted
$`e^{\pi /4kr_C}\mathrm{\Gamma }\left(1+{\displaystyle \frac{i}{2kr_C}}\right)=e^{i\sigma _0\left(k\right)}C_0(k),\sigma _0(k)`$ $`=`$ $`\mathrm{arg}\mathrm{\Gamma }\left(1+{\displaystyle \frac{i}{2kr_C}}\right),`$ (6.12)
where $`\sigma _0(k)`$ is a pure Coulomb phase shift.
Now, let us rewrite the denominator of the amplitude Eq.(6) in the equivalent form
$`\{\mathrm{cos}\sigma _0(k)[1a(a+1){\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}M_\mathrm{N}^2+{\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}(b^2a(a+1))k^2]`$
$`\mathrm{sin}\sigma _0(k){\displaystyle \frac{C_{\mathrm{NN}}M_\mathrm{N}}{2\pi }}kC_0^2(k)\}i\{\mathrm{cos}\sigma _0(k){\displaystyle \frac{G_{\pi \mathrm{NN}}M_\mathrm{N}}{2\pi }}kC_0^2(k)`$
$`+\mathrm{sin}\sigma _0(k)[1a(a+1){\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}M_\mathrm{N}^2+{\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}(b^2a(a+1))k^2]\}=`$
$`={\displaystyle \frac{1}{๐ต}}\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)+ia_{\mathrm{pp}}^\mathrm{e}kC_0^2(k)\right],`$ (6.13)
where we have denoted
$`{\displaystyle \frac{1}{๐ต}}\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)\right]=\mathrm{sin}\sigma _0(k){\displaystyle \frac{C_{\mathrm{NN}}M_\mathrm{N}}{2\pi }}kC_0^2(k)`$
$`+\mathrm{cos}\sigma _0(k)\left[1a(a+1){\displaystyle \frac{G_{\pi \mathrm{NN}}}{2\pi ^2}}M_\mathrm{N}^2+{\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}(b^2a(a+1))k^2\right],`$
$`{\displaystyle \frac{1}{๐ต}}a_{\mathrm{pp}}^\mathrm{e}kC_0^2(k)=\mathrm{cos}\sigma _0(k){\displaystyle \frac{C_{\mathrm{NN}}M_\mathrm{N}}{2\pi }}kC_0^2(k)`$
$`+\mathrm{sin}\sigma _0(k)\left[1a(a+1){\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}M_\mathrm{N}^2+{\displaystyle \frac{C_{\mathrm{NN}}}{2\pi ^2}}(b^2a(a+1))k^2\right].`$ (6.14)
Here $`๐ต`$ is a constant which we would remove by the renormalization of the wave functions of the protons, $`a_{\mathrm{pp}}^\mathrm{e}=(7.8196\pm 0.0026)\mathrm{fm}`$ and $`r_{\mathrm{pp}}^\mathrm{e}=2.790\pm 0.014\mathrm{fm}`$ are the Sโwave scattering length and the effective range of pp scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state with the Coulomb repulsion, and $`h(2kr_C)`$ is defined by
$`h(2kr_C)=\gamma +\mathrm{}n(2kr_C)+{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n(1+4n^2k^2r_C^2)}}.`$ (6.15)
The validity of the relations Eq.(6) assumes the dependence of parameters $`a`$ and $`b`$ on the relative momentum $`k`$.
After the changes Eq.(6) and Eq.(6) the amplitude Eq.(6) takes the form
$`i(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)=G_\mathrm{V}g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}G_\mathrm{V}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}})\gamma ^\mu (1\gamma ^5)v(k_{\mathrm{e}^+})]`$
$`\times {\displaystyle \frac{C_0(k)}{1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)+ia_{\mathrm{pp}}^\mathrm{e}kC_0^2(k)}}๐ต[\overline{u^c}(p_2)\gamma ^5u(p_1)].`$ (6.16)
Renormalizing the wave functions of the protons $`\sqrt{๐ต}u(p_2)u(p_2)`$ and $`\sqrt{๐ต}u(p_1)u(p_1)`$ we obtain the amplitude of the solar proton burning
$`i(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)=g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}G_\mathrm{V}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}})\gamma ^\mu (1\gamma ^5)v(k_{\mathrm{e}^+})]`$
$`\times {\displaystyle \frac{C_0(k)}{1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)+ia_{\mathrm{pp}}^\mathrm{e}kC_0^2(k)}}[\overline{u^c}(p_2)\gamma ^5u(p_1)]F_\mathrm{D}(k^2),`$ (6.17)
where $`F_\mathrm{D}(k^2)`$ is given by Eq.(2.16) and describes the spatial smearing of the deuteron coupled to the pp pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state.
The real part of the denominator of the amplitude Eq.(6) is in complete agreement with a phenomenological relation
$`\mathrm{ctg}\delta _{\mathrm{pp}}^\mathrm{e}(k)={\displaystyle \frac{1}{C_0^2(k)k}}\left[{\displaystyle \frac{1}{a_{\mathrm{pp}}^\mathrm{e}}}+{\displaystyle \frac{1}{2}}r_{\mathrm{pp}}^\mathrm{e}k^2{\displaystyle \frac{1}{r_\mathrm{C}}}h(2kr_\mathrm{C})\right],`$ (6.18)
describing the phase shift $`\delta _{\mathrm{pp}}^\mathrm{e}(k)`$ of lowโenergy elastic pp scattering in terms of the Sโwave scattering length $`a_{\mathrm{pp}}^\mathrm{e}`$ and the effective range $`r_{\mathrm{pp}}^\mathrm{e}`$. Thus, we argue that the contribution of lowโenergy elastic pp scattering to the amplitude of the solar proton burning is described in agreement with lowโenergy nuclear phenomenology in terms of the Sโwave scattering length $`a_{\mathrm{pp}}^\mathrm{e}`$ and the effective range $`r_{\mathrm{pp}}^\mathrm{e}`$ taken from the experimental data .
## 7 The astrophysical factor for the solar proton burning
The amplitude Eq.(6) squared, averaged over polarizations of the protons and summed over polarizations of final particles reads
$`\overline{|(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)|^2}=G_\mathrm{V}^2g_\mathrm{A}^2M_\mathrm{N}^6C_{\mathrm{NN}}^2{\displaystyle \frac{54Q_\mathrm{D}}{\pi ^2}}F_\mathrm{D}^2(k^2)`$ (7.1)
$`\times {\displaystyle \frac{C_0^2(k)}{\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)\right]^2+(a_{\mathrm{pp}}^\mathrm{e})^2k^2C_0^4(k)}}\left(E_{\mathrm{e}^+}E_{\nu _\mathrm{e}}{\displaystyle \frac{1}{3}}\stackrel{}{k}_{\mathrm{e}^+}\stackrel{}{k}_{\nu _\mathrm{e}}\right),`$
where $`m_\mathrm{e}=0.511\mathrm{MeV}`$ is the positron mass, and we have used the relation $`g_\mathrm{V}^2=2\pi ^2Q_\mathrm{D}M_\mathrm{N}^2`$.
The cross section for the reaction p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ is defined by
$`\sigma ^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(T_{\mathrm{pp}})`$ $`=`$ $`{\displaystyle \frac{1}{v}}{\displaystyle \frac{1}{4E_1E_2}}{\displaystyle \overline{|(\mathrm{p}+\mathrm{p}\mathrm{D}+\mathrm{e}^++\nu _e)|^2}}`$ (7.2)
$`\times `$ $`(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\mathrm{}}p_1p_2){\displaystyle \frac{d^3k_\mathrm{D}}{(2\pi )^32E_\mathrm{D}}}{\displaystyle \frac{d^3k_{\mathrm{e}^+}}{(2\pi )^32E_{\mathrm{e}^+}}}{\displaystyle \frac{d^3k_{\nu _\mathrm{e}}}{(2\pi )^32E_{\nu _\mathrm{e}}}},`$
where $`v`$ is a relative velocity of the pp pair and $`k_{\mathrm{}}=k_{\mathrm{e}^+}+k_{\nu _\mathrm{e}}`$ is a 4โmomentum of the leptonic pair.
The integration over the phase volume of the final $`\mathrm{De}^+\nu _\mathrm{e}`$ state we perform in the nonโrelativistic limit
$`{\displaystyle \frac{d^3k_\mathrm{D}}{(2\pi )^32E_\mathrm{D}}\frac{d^3k_{\mathrm{e}^+}}{(2\pi )^32E_{\mathrm{e}^+}}\frac{d^3k_{\nu _\mathrm{e}}}{(2\pi )^32E_{\nu _\mathrm{e}}}(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\mathrm{}}p_1p_2)\left(E_{\mathrm{e}^+}E_{\nu _\mathrm{e}}\frac{1}{3}\stackrel{}{k}_{\mathrm{e}^+}\stackrel{}{k}_{\nu _\mathrm{e}}\right)}`$
$`={\displaystyle \frac{1}{32\pi ^3M_\mathrm{N}}}{\displaystyle _{m_\mathrm{e}}^{W+T_{\mathrm{pp}}}}\sqrt{E_{\mathrm{e}^+}^2m_\mathrm{e}^2}E_{\mathrm{e}^+}(W+T_{\mathrm{pp}}E_{\mathrm{e}^+})^2๐E_{\mathrm{e}^+}={\displaystyle \frac{(W+T_{\mathrm{pp}})^5}{960\pi ^3M_\mathrm{N}}}f(\xi ),`$ (7.3)
where $`W=\epsilon _\mathrm{D}(M_\mathrm{n}M_\mathrm{p})=(2.2251.293)\mathrm{MeV}=0.932\mathrm{MeV}`$ and $`\xi =m_\mathrm{e}/(W+T_{\mathrm{pp}})`$. The function $`f(\xi )`$ is defined by the integral
$`f(\xi )`$ $`=`$ $`30{\displaystyle _\xi ^1}\sqrt{x^2\xi ^2}x(1x)^2๐x=(1{\displaystyle \frac{9}{2}}\xi ^24\xi ^4)\sqrt{1\xi ^2}`$ (7.4)
$`+{\displaystyle \frac{15}{2}}\xi ^4\mathrm{}n\left({\displaystyle \frac{1+\sqrt{1\xi ^2}}{\xi }}\right)|_{T_{\mathrm{pp}}=0}=0.222`$
and normalized to unity at $`\xi =0`$.
Thus, the cross section for the solar proton burning is given by
$`\sigma ^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(T_{\mathrm{pp}})={\displaystyle \frac{e^{\pi /r_C\sqrt{M_\mathrm{N}T_{\mathrm{pp}}}}}{v^2}}\alpha {\displaystyle \frac{9g_\mathrm{A}^2G_\mathrm{V}^2Q_\mathrm{D}M_\mathrm{N}^3}{320\pi ^4}}C_{\mathrm{NN}}^2(W+T_{\mathrm{pp}})^5f\left({\displaystyle \frac{m_\mathrm{e}}{W+T_{\mathrm{pp}}}}\right)`$
$`\times {\displaystyle \frac{F_\mathrm{D}^2(M_\mathrm{N}T_{\mathrm{pp}})}{\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}M_\mathrm{N}T_{\mathrm{pp}}+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2r_C\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})\right]^2+(a_{\mathrm{pp}}^\mathrm{e})^2M_\mathrm{N}T_{\mathrm{pp}}C_0^4(\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})}}`$
$`\times {\displaystyle \frac{1}{1e^{\pi /r_C\sqrt{M_\mathrm{N}T_{\mathrm{pp}}}}}}={\displaystyle \frac{S_{\mathrm{pp}}(T_{\mathrm{pp}})}{T_{\mathrm{pp}}}}e^{\pi /r_C\sqrt{M_\mathrm{N}T_{\mathrm{pp}}}}.`$ (7.5)
The astrophysical factor $`S_{\mathrm{pp}}(T_{\mathrm{pp}})`$ reads
$`S_{\mathrm{pp}}(T_{\mathrm{pp}})=\alpha {\displaystyle \frac{9g_\mathrm{A}^2G_\mathrm{V}^2Q_\mathrm{D}M_\mathrm{N}^4}{1280\pi ^4}}{\displaystyle \frac{C_{\mathrm{NN}}^2(W+T_{\mathrm{pp}})^5}{1e^{\pi /r_C\sqrt{M_\mathrm{N}T_{\mathrm{pp}}}}}}f\left({\displaystyle \frac{m_\mathrm{e}}{W+T_{\mathrm{pp}}}}\right)`$
$`\times {\displaystyle \frac{F_\mathrm{D}^2(M_\mathrm{N}T_{\mathrm{pp}})}{\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}M_\mathrm{N}T_{\mathrm{pp}}+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2r_C\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})\right]^2+(a_{\mathrm{pp}}^\mathrm{e})^2M_\mathrm{N}T_{\mathrm{pp}}C_0^4(\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})}}.`$ (7.6)
At zero kinetic energy of the relative movement of the protons $`T_{\mathrm{pp}}=0`$ the astrophysical factor $`S_{\mathrm{pp}}(0)`$ is given by
$`S_{\mathrm{pp}}(0)=\alpha {\displaystyle \frac{9g_\mathrm{A}^2G_\mathrm{V}^2Q_\mathrm{D}M_\mathrm{N}^4}{1280\pi ^4}}C_{\mathrm{NN}}^2W^5f\left({\displaystyle \frac{m_\mathrm{e}}{W}}\right)=4.08\times 10^{25}\mathrm{MeV}\mathrm{b}.`$ (7.7)
The value $`S_{\mathrm{pp}}(0)=4.08\times 10^{25}\mathrm{MeV}\mathrm{b}`$ agrees well with the recommended value $`S_{\mathrm{pp}}(0)=4.00\times 10^{25}\mathrm{MeV}\mathrm{b}`$ .
Unlike the astrophysical factor obtained by Kamionkowski and Bahcall the astrophysical factor given by Eq.(7.7) does not depend explicitly on the Sโwave scattering length of lowโenergy elastic pp scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state. This is due to the normalization of the wave function of the pp pair. After the summation of an infinite series and by using the relation Eq.(6.18) we obtain the wave function of the pp pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state in the form
$`\psi _{\mathrm{pp}}(k)=e^{i\delta _{\mathrm{pp}}^\mathrm{e}\left(k\right)}{\displaystyle \frac{\mathrm{sin}\delta _{\mathrm{pp}}^\mathrm{e}(k)}{a_{\mathrm{pp}}^\mathrm{e}kC_0(k)}},`$ (7.8)
that corresponds the normalization of the wave function of the relative movement of the pp pair used by Schiavilla et al. . For the more detailed discussion of this problem we relegate readers to the paper by Schiavilla et al. <sup>3</sup><sup>3</sup>3See the last paragraph of Sect. 3 and the first paragraph of Sect. 5 of Ref...
## 8 The reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\mathrm{e}^{}`$ \+ p + p
The evaluation of the amplitude of the reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\mathrm{e}^{}`$ \+ p + p is analogous to that of the amplitude of the solar proton burning. The result reads
$`i(\nu _e+\mathrm{D}\mathrm{e}^{}+\mathrm{p}+\mathrm{p})=g_\mathrm{A}M_\mathrm{N}{\displaystyle \frac{G_\mathrm{V}}{\sqrt{2}}}{\displaystyle \frac{3g_\mathrm{V}}{2\pi ^2}}C_{\mathrm{NN}}e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_\mathrm{e}^{})\gamma ^\mu (1\gamma ^5)u(k_{\nu _\mathrm{e}})]`$
$`\times {\displaystyle \frac{C_0(k)}{1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)+ia_{\mathrm{pp}}^\mathrm{e}kC_0^2(k)}}[\overline{u}(p_2)\gamma ^5u^c(p_1)]F_\mathrm{D}(k^2).`$ (8.1)
The amplitude Eq.(8) squared, averaged over polarizations of the deuteron and summed over polarizations of the final particles reads
$`\overline{|(\nu _\mathrm{e}+\mathrm{D}\mathrm{e}^{}+\mathrm{p}+\mathrm{p})|^2}=S_{\mathrm{pp}}(0){\displaystyle \frac{2^{12}5\pi ^2}{\mathrm{\Omega }_{\mathrm{De}^+\nu _\mathrm{e}}}}{\displaystyle \frac{r_\mathrm{C}M_\mathrm{N}^3}{m_\mathrm{e}^5}}F_\mathrm{D}^2(k^2)F_+(Z,E_\mathrm{e}^{})`$
$`\times {\displaystyle \frac{C_0^2(k)}{\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_\mathrm{C}}}h(2kr_\mathrm{C})\right]^2+(a_{\mathrm{pp}}^\mathrm{e})^2k^2C_0^4(k)}}\left(E_\mathrm{e}^{}E_{\nu _\mathrm{e}}{\displaystyle \frac{1}{3}}\stackrel{}{k}_\mathrm{e}^{}\stackrel{}{k}_{\nu _\mathrm{e}}\right).`$ (8.2)
where $`F_+(Z,E_\mathrm{e}^{})`$ is the Fermi function describing the Coulomb interaction of the electron with the nuclear system having a charge $`Z`$. In the case of the reaction $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p we have $`Z=2`$. At $`\alpha ^2Z^21`$ the Fermi function $`F_+(Z,E_\mathrm{e}^{})`$ reads
$`{}_{+}{}^{}F(Z,E_\mathrm{e}^{})={\displaystyle \frac{2\pi \eta _\mathrm{e}^{}}{1e^{2\pi \eta _\mathrm{e}^{}}}},`$ (8.3)
where $`\eta _\mathrm{e}^{}=Z\alpha /v_\mathrm{e}^{}=Z\alpha E_\mathrm{e}^{}/\sqrt{E_\mathrm{e}^{}^2m_\mathrm{e}^{}^2}`$ and $`v_\mathrm{e}^{}`$ is a velocity of the electron.
For the evaluation of the r.h.s. of Eq.(8) we have also used the expression for the astrophysical factor
$`S_{\mathrm{pp}}(0)={\displaystyle \frac{9g_\mathrm{A}^2G_\mathrm{V}^2Q_\mathrm{D}M_\mathrm{N}^3}{1280\pi ^4r_\mathrm{C}}}C_{\mathrm{NN}}^2m_\mathrm{e}^5\mathrm{\Omega }_{\mathrm{De}^+\nu _\mathrm{e}},`$ (8.4)
where $`\mathrm{\Omega }_{\mathrm{De}^+\nu _\mathrm{e}}=(W/m_\mathrm{e})^5f(m_\mathrm{e}/W)=4.481`$ at $`W=0.932\mathrm{MeV}`$. The function $`f(m_\mathrm{e}/W)`$ is defined by Eq.(7.4).
In the rest frame of the deuteron the cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\mathrm{e}^{}`$ \+ p + p is defined by
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}})={\displaystyle \frac{1}{4M_\mathrm{D}E_{\nu _\mathrm{e}}}}{\displaystyle \overline{|(\nu _\mathrm{e}+\mathrm{D}\mathrm{e}^{}+\mathrm{p}+\mathrm{p})|^2}}`$
$`{\displaystyle \frac{1}{2}}(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\nu _\mathrm{e}}p_1p_2k_\mathrm{e}^{}){\displaystyle \frac{d^3p_1}{(2\pi )^32E_1}}{\displaystyle \frac{d^3p_2}{(2\pi )^32E_2}}{\displaystyle \frac{d^3k_\mathrm{e}^{}}{(2\pi )^32E_\mathrm{e}^{}}},`$ (8.5)
where $`E_{\nu _\mathrm{e}}`$, $`E_1`$, $`E_2`$ and $`E_\mathrm{e}^{}`$ are the energies of the neutrino, the protons and the electron. The integration over the phase volume of the ($`\mathrm{ppe}^{}`$) state we perform in the nonโrelativistic limit and in the rest frame of the deuteron,
$`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^3p_1}{(2\pi )^32E_1}\frac{d^3p_2}{(2\pi )^32E_2}\frac{d^3k_\mathrm{e}}{(2\pi )^32E_\mathrm{e}^{}}(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\nu _\mathrm{e}}p_1p_2k_\mathrm{e}^{})}`$
$`{\displaystyle \frac{C_0^2(\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})F_\mathrm{D}^2(M_\mathrm{N}T_{\mathrm{pp}})F_+(Z,E_\mathrm{e}^{})}{\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}M_\mathrm{N}T_{\mathrm{pp}}+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_\mathrm{C}}}h(2r_\mathrm{C}\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})\right]^2+(a_{\mathrm{pp}}^\mathrm{e})^2M_\mathrm{N}T_{\mathrm{pp}}C_0^4(\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})}}`$
$`\left(E_\mathrm{e}^{}E_{\nu _\mathrm{e}}{\displaystyle \frac{1}{3}}\stackrel{}{k}_\mathrm{e}^{}\stackrel{}{k}_{\nu _\mathrm{e}}\right)={\displaystyle \frac{E_{\nu _\mathrm{e}}M_\mathrm{N}^3}{128\pi ^3}}\left({\displaystyle \frac{E_{\mathrm{th}}}{M_\mathrm{N}}}\right)^{7/2}\left({\displaystyle \frac{2m_\mathrm{e}}{E_{\mathrm{th}}}}\right)^{3/2}{\displaystyle \frac{1}{E_{\mathrm{th}}^2}}`$
$`{\displaystyle ๐T_\mathrm{e}^{}๐T_{\mathrm{pp}}\delta (E_{\nu _\mathrm{e}}E_{\mathrm{th}}T_\mathrm{e}^{}T_{\mathrm{pp}})\sqrt{T_\mathrm{e}^{}T_{\mathrm{pp}}}\left(1+\frac{T_\mathrm{e}^{}}{m_\mathrm{e}}\right)\sqrt{1+\frac{T_\mathrm{e}^{}}{2m_\mathrm{e}}}}`$
$`{\displaystyle \frac{C_0^2(\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})F_\mathrm{D}^2(M_\mathrm{N}T_{\mathrm{pp}})F_+(Z,E_\mathrm{e}^{})}{\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}M_\mathrm{N}T_{\mathrm{pp}}+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_\mathrm{C}}}h(2r_\mathrm{C}\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})\right]^2+(a_{\mathrm{pp}}^\mathrm{e})^2M_\mathrm{N}T_{\mathrm{pp}}C_0^4(\sqrt{M_\mathrm{N}T_{\mathrm{pp}}})}}`$
$`={\displaystyle \frac{E_{\nu _\mathrm{e}}M_\mathrm{N}^3}{128\pi ^3}}\left({\displaystyle \frac{E_{\mathrm{th}}}{M_\mathrm{N}}}\right)^{7/2}\left({\displaystyle \frac{2m_\mathrm{e}}{E_{\mathrm{th}}}}\right)^{3/2}(y1)^2\mathrm{\Omega }_{\mathrm{ppe}^{}}(y),`$ (8.6)
where $`T_\mathrm{e}^{}`$ is the kinetic energy of the electron, $`E_{\mathrm{th}}`$ is the neutrino energy threshold: $`E_{\mathrm{th}}=\epsilon _\mathrm{D}+m_\mathrm{e}(M_\mathrm{n}M_\mathrm{p})=(2.225+0.5111.293)\mathrm{MeV}=1.443\mathrm{MeV}`$. The function $`\mathrm{\Omega }_{\mathrm{ppe}^{}}(y)`$, where $`y=E_{\nu _\mathrm{e}}/E_{\mathrm{th}}`$ and $`E_{\mathrm{th}}=\epsilon _\mathrm{D}(M_\mathrm{n}M_\mathrm{p})+m_\mathrm{e}=1.443\mathrm{MeV}`$, is determined by the integral
$`\mathrm{\Omega }_{\mathrm{ppe}^{}}(y)={\displaystyle \underset{0}{\overset{1}{}}}๐x\sqrt{x(1x)}\left(1+{\displaystyle \frac{E_{\mathrm{th}}}{m_\mathrm{e}}}(y1)(1x)\right)\sqrt{1+{\displaystyle \frac{E_{\mathrm{th}}}{2m_\mathrm{e}}}(y1)(1x)}`$
$`C_0^2(\sqrt{M_\mathrm{N}E_{\mathrm{th}}(y1)x})F_\mathrm{D}^2(M_\mathrm{N}E_{\mathrm{th}}(y1)x)F_+(Z,m_\mathrm{e}+E_{\mathrm{th}}(y1)(1x))`$
$`\{[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}M_\mathrm{N}E_{\mathrm{th}}(y1)x+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_\mathrm{C}}}h(2r_\mathrm{C}\sqrt{M_\mathrm{N}E_{\mathrm{th}}(y1)x})]^2`$
$`+(a_{\mathrm{pp}}^\mathrm{e})^2M_\mathrm{N}E_{\mathrm{th}}(y1)xC_0^4(\sqrt{M_\mathrm{N}E_{\mathrm{th}}(y1)x})\}^1,`$ (8.7)
where we have changed the variable $`T_{\mathrm{pp}}=(E_{\nu _\mathrm{e}}E_{\mathrm{th}})x`$.
The cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\mathrm{e}^{}`$ \+ p + p is defined by
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}})`$ $`=`$ $`S_{\mathrm{pp}}(0){\displaystyle \frac{640r_\mathrm{C}}{\pi \mathrm{\Omega }_{\mathrm{De}^+\nu _\mathrm{e}}}}\left({\displaystyle \frac{M_\mathrm{N}}{E_{\mathrm{th}}}}\right)^{3/2}\left({\displaystyle \frac{E_{\mathrm{th}}}{2m_\mathrm{e}}}\right)^{7/2}(y1)^2\mathrm{\Omega }_{\mathrm{ppe}^{}}(y)=`$ (8.8)
$`=`$ $`3.69\times 10^5S_{\mathrm{pp}}(0)(y1)^2\mathrm{\Omega }_{\mathrm{ppe}^{}}(y),`$
where $`S_{\mathrm{pp}}(0)`$ is measured in $`\mathrm{MeV}\mathrm{cm}^2`$. For $`S_{\mathrm{pp}}(0)=4.08\times 10^{49}\mathrm{MeV}\mathrm{cm}^2`$ Eq.(7.7) the cross section $`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}})`$ reads
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}})=1.50(y1)^2\mathrm{\Omega }_{\mathrm{ppe}^{}}(y)\mathrm{\hspace{0.17em}10}^{43}\mathrm{cm}^2.`$ (8.9)
Recently the calculation of the cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p has been carried in Ref. within the PMA and tabulated for the neutrino energies ranging over the region from threshold up to 170$`\mathrm{MeV}`$. Since our result is valid for much lower neutrino energies, we give the numerical values of the cross section only for the energies from the region $`4\mathrm{MeV}E_{\nu _\mathrm{e}}10\mathrm{MeV}`$:
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}}=4\mathrm{MeV})`$ $`=`$ $`2.46(1.577)\times 10^{43}\mathrm{cm}^2,`$
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}}=6\mathrm{MeV})`$ $`=`$ $`10.04(6.239)\times 10^{43}\mathrm{cm}^2,`$
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}}=8\mathrm{MeV})`$ $`=`$ $`2.37(1.463)\times 10^{42}\mathrm{cm}^2,`$
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}}=10\mathrm{MeV})`$ $`=`$ $`4.39(2.708)\times 10^{43}\mathrm{cm}^2.`$ (8.10)
The data in parentheses are taken from Table 1 of Ref. . Thus, on the average the cross section $`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}})`$ calculated in the NNJL model by a factor of 1.6 is larger compared with the PMA ones.
Since the amplitude of the reaction $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p \+ p is completely defined by the amplitude of the solar proton burning p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ that is described in the NNJL model in agreement with other theoretical approaches, our prediction for the cross section for the neutrino disintegration of the deuteron Eq.(8) is a challenge to the experiments planned at SNO .
In order to compare the cross section for the neutrino disintegration of the deuteron $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p with experimental data planned to be obtained by SNO Collaboration we should average it over the $`{}_{}{}^{8}\mathrm{B}`$ solar neutrino energy spectrum produced by the $`\beta `$ decay $`{}_{}{}^{8}\mathrm{B}{}_{}{}^{8}\mathrm{Be}_{}^{}`$ \+ e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ in the solar core. Using the $`{}_{}{}^{8}\mathrm{B}`$ solar neutrino energy spectrum and integrating over the region $`5\mathrm{MeV}E_{\nu _\mathrm{e}}15\mathrm{MeV}`$ , where the lower bound is related to the experimental threshold of experiments at SNO and the upper one is defined by the kinematics of the decay $`{}_{}{}^{8}\mathrm{B}{}_{}{}^{8}\mathrm{Be}_{}^{}`$ \+ e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, we get
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}})_{\mathrm{\Phi }({}_{}{}^{8}\mathrm{B})}=2.62\times 10^{42}\mathrm{cm}^2.`$ (8.11)
For the comparison of the theoretical cross section with the experimental one measured at SNO one should take into account a possible decrease of the experimental value in the case of the existence of neutrino flavour oscillations .
## 9 The astrophysical factor for the pep process
In the NNJL model the amplitude of the reaction p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$ or the pepโprocess is related to the amplitude of the solar proton burning p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ Eq.(6) as well as to the amplitude of the neutrino disintegration of the deuteron $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p. Indeed, the effective Lagrangians $`_{\mathrm{eff}}^{\mathrm{pe}^{}\mathrm{p}\mathrm{D}\nu _\mathrm{e}}(x)`$, $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ and $`_{\mathrm{eff}}^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(x)`$ are defined by the anomaly of the oneโnucleon loop triangle $`AAV`$โdiagrams with two axialโvector $`(A)`$ and one vector $`(V)`$ vertices (see Appendix). The amplitude of the pepโprocess reads
$`i(\mathrm{p}+\mathrm{e}^{}+\mathrm{p}\mathrm{D}+\nu _e)=g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}G_\mathrm{V}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}e_\mu ^{}(k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}})\gamma ^\mu (1\gamma ^5)u(k_\mathrm{e}^{})]`$
$`\times {\displaystyle \frac{C_0(k)}{1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)+ia_{\mathrm{pp}}^\mathrm{e}kC_0^2(k)}}[\overline{u^c}(p_2)\gamma ^5u(p_1)]F_\mathrm{D}(k^2).`$ (9.1)
The amplitude Eq.(9) squared, averaged and summed over polarizations of the interacting particles is defined by
$`\overline{|(\mathrm{p}+\mathrm{e}^{}+\mathrm{p}\mathrm{D}+\nu _e)|^2}=g_\mathrm{A}^2M_\mathrm{N}^6C_{\mathrm{NN}}^2G_\mathrm{V}^2{\displaystyle \frac{27Q_\mathrm{D}}{\pi ^2}}F_\mathrm{D}^2(k^2)F_+(Z,E_\mathrm{e}^{})`$ (9.2)
$`\times {\displaystyle \frac{C_0^2(k)}{\left[1{\displaystyle \frac{1}{2}}a_{\mathrm{pp}}^\mathrm{e}r_{\mathrm{pp}}^\mathrm{e}k^2+{\displaystyle \frac{a_{\mathrm{pp}}^\mathrm{e}}{r_C}}h(2kr_C)\right]^2+(a_{\mathrm{pp}}^\mathrm{e})^2k^2C_0^4(k)}}\left(E_{\mathrm{e}^+}E_{\nu _\mathrm{e}}{\displaystyle \frac{1}{3}}\stackrel{}{k}_{\mathrm{e}^+}\stackrel{}{k}_{\nu _\mathrm{e}}\right),`$
where $`F_+(Z,E_\mathrm{e}^{})`$ is the Fermi function given by Eq.(8.3).
At low energies the cross section $`\sigma ^{\mathrm{pe}^{}\mathrm{p}\mathrm{D}\nu _\mathrm{e}}(T_{\mathrm{pp}})`$ for the pepโprocess can be determined as follows
$`\sigma ^{\mathrm{pe}^{}\mathrm{p}\mathrm{D}\nu _\mathrm{e}}(T_{\mathrm{pp}})={\displaystyle \frac{1}{v}}{\displaystyle \frac{1}{4M_\mathrm{N}^2}}{\displaystyle \frac{d^3k_\mathrm{e}^{}}{(2\pi )^32E_\mathrm{e}^{}}gn(\stackrel{}{k}_\mathrm{e}^{})\overline{|(\mathrm{p}+\mathrm{e}^{}+\mathrm{p}\mathrm{D}+\nu _\mathrm{e})|^2}}`$
$`(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\nu _\mathrm{e}}p_1p_2k_\mathrm{e}^{}){\displaystyle \frac{d^3k_\mathrm{D}}{(2\pi )^32M_\mathrm{D}}}{\displaystyle \frac{d^3k_{\nu _\mathrm{e}}}{(2\pi )^32E_{\nu _\mathrm{e}}}},`$ (9.3)
where $`g=2`$ is the number of the electron spin states and $`v`$ is a relative velocity of the pp pair. The electron distribution function $`n(\stackrel{}{k}_\mathrm{e}^{})`$ can be taken in the form
$`n(\stackrel{}{k}_\mathrm{e}^{})=e^{\overline{\nu }T_\mathrm{e}^{}/kT_c},`$ (9.4)
where $`k=8.617\times 10^{11}\mathrm{MeV}\mathrm{K}^1`$, $`T_c`$ is a temperature of the core of the Sun. The distribution function $`n(\stackrel{}{k}_\mathrm{e}^{})`$ is normalized by the condition
$`g{\displaystyle \frac{d^3k_\mathrm{e}^{}}{(2\pi )^3}n(\stackrel{}{k}_\mathrm{e}^{})}=n_\mathrm{e}^{},`$ (9.5)
where $`n_\mathrm{e}^{}`$ is the electron number density. From the normalization condition Eq.(9.5) we derive
$`e^{\overline{\nu }}={\displaystyle \frac{4\pi ^3n_\mathrm{e}^{}}{(2\pi m_\mathrm{e}kT_c)^{3/2}}}.`$ (9.6)
The astrophysical factor $`S_{\mathrm{pep}}(0)`$ is then defined by
$`S_{\mathrm{pep}}(0)=S_{\mathrm{pp}}(0){\displaystyle \frac{15}{2\pi }}{\displaystyle \frac{1}{\mathrm{\Omega }_{\mathrm{De}^+\nu _\mathrm{e}}}}{\displaystyle \frac{1}{m_\mathrm{e}^3}}\left({\displaystyle \frac{E_{\mathrm{th}}}{m_\mathrm{e}}}\right)^2e^{\overline{\nu }}{\displaystyle d^3k_\mathrm{e}^{}e^{T_\mathrm{e}^{}/kT_c}F_+(Z,E_\mathrm{e}^{})}.`$ (9.7)
For the ratio $`S_{\mathrm{pep}}(0)/S_{\mathrm{pp}}(0)`$ we obtain
$`{\displaystyle \frac{S_{\mathrm{pep}}(0)}{S_{\mathrm{pp}}(0)}}={\displaystyle \frac{2^{3/2}\pi ^{5/2}}{f_{\mathrm{pp}}(0)}}\left({\displaystyle \frac{\alpha Zn_\mathrm{e}^{}}{m_\mathrm{e}^3}}\right)\left({\displaystyle \frac{E_{\mathrm{th}}}{m_\mathrm{e}}}\right)^2\sqrt{{\displaystyle \frac{m_\mathrm{e}}{kT_c}}}I\left(Z\sqrt{{\displaystyle \frac{2m_\mathrm{e}}{kT_c}}}\right).`$ (9.8)
We have set $`f_{\mathrm{pp}}(0)=\mathrm{\Omega }_{\mathrm{De}^+\nu _\mathrm{e}}/30=0.149`$ and the function $`I(x)`$ introduced by Bahcall and May reads
$`I(x)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{due^u}{1e^{\pi \alpha x/\sqrt{u}}}}.`$ (9.9)
The relation between the astrophysical factors $`S_{\mathrm{pep}}(0)`$ and $`S_{\mathrm{pp}}(0)`$ given by Eq.(9.8) is in complete agreement with that obtained by Bahcall and May .
By virtue of the direct relation between the amplitudes of the pepโprocess and the reaction for the neutrino disintegration of the deuteron $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p that we have in the NNJL model the agreement with the result obtained by Bahcall and May is on favour of our predictions for the cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p.
## 10 The reaction $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n
The effective Lagrangian $`_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)`$ of the lowโenergy nuclear transition $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n we evaluate by analogy with $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\overline{\nu }_\mathrm{e}}(x)`$ (see Eq.(A.27) of Appendix) through the oneโnucleon loop exchanges and at leading order in the large $`N_C`$ expansion. The effective Lagrangian $`_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)`$ is defined by the anomaly of the oneโnucleon loop triangle $`AAV`$โdiagram as well as the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$. The result reads
$`_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)=ig_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}{\displaystyle \frac{G_\mathrm{V}}{\sqrt{2}}}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}D_\mu (x)[\overline{n}(x)\gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\mu (1\gamma ^5)\psi _\mathrm{e}(x)].`$ (10.1)
The amplitude of the reaction $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n \+ n we obtain in the form
$`i(\overline{\nu }_\mathrm{e}+\mathrm{D}\mathrm{e}^++\mathrm{n}+\mathrm{n})`$ $`=`$ $`g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}{\displaystyle \frac{G_\mathrm{V}}{\sqrt{2}}}{\displaystyle \frac{3g_\mathrm{V}}{2\pi ^2}}{\displaystyle \frac{1}{1{\displaystyle \frac{1}{2}}r_{\mathrm{nn}}a_{\mathrm{nn}}k^2+ia_{\mathrm{nn}}k}}F_\mathrm{D}(k^2)`$ (10.2)
$`\times e_\mu (k_\mathrm{D})[\overline{v}(k_{\overline{\nu }_\mathrm{e}})\gamma ^\mu (1\gamma ^5)v(k_{\mathrm{e}^+})][\overline{u}(p_2)\gamma ^5u^c(p_1)],`$
where the form factor $`F_\mathrm{D}(k^2)`$ describes a spatial smearing of the deuteron Eq.(2.16). The factor $`1/(1\frac{1}{2}r_{\mathrm{nn}}a_{\mathrm{nn}}k^2+ia_{\mathrm{nn}}k)`$ gives the contribution of lowโenergy nuclear forces to the relative movement of the nn pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state. The result is obtained by the summation of an infinite series of the oneโneutron bubbles evaluated at leading order in the large $`N_C`$ expansion. Since we work in the isotopical limit, we set $`a_{\mathrm{nn}}=a_{\mathrm{np}}=23.75\mathrm{fm}`$ and $`r_{\mathrm{nn}}=r_{\mathrm{np}}=2.75\mathrm{fm}`$. The recent experimental values of the Sโwave scattering length and the effective range of lowโenergy elastic nn scattering are equal to $`a_{\mathrm{nn}}=(18.8\pm 0.3)\mathrm{fm}`$ and $`r_{\mathrm{nn}}=(2.75\pm 0.11)\mathrm{fm}`$ .
The amplitude Eq. (10.2), squared, averaged over polarizations of the deuteron and summed over polarizations of the final particles, reads
$`\overline{|(\overline{\nu }_\mathrm{e}+\mathrm{D}\mathrm{e}^++\mathrm{n}+\mathrm{n})|^2}={\displaystyle \frac{144}{\pi ^2}}{\displaystyle \frac{Q_\mathrm{D}g_\mathrm{A}^2G_\mathrm{V}^2C_{\mathrm{NN}}^2M_\mathrm{N}^6F_\mathrm{D}^2(k^2)}{\left(1{\displaystyle \frac{1}{2}}r_{\mathrm{nn}}a_{\mathrm{nn}}k^2\right)^2+a_{\mathrm{nn}}^2k^2}}\left(E_{\mathrm{e}^+}E_{\overline{\nu }_\mathrm{e}}{\displaystyle \frac{1}{3}}\stackrel{}{k}_{\mathrm{e}^+}\stackrel{}{k}_{\overline{\nu }_\mathrm{e}}\right).`$ (10.3)
The cross section for the reaction $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n is defined by
$`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(E_{\overline{\nu }_\mathrm{e}})={\displaystyle \frac{1}{4E_\mathrm{D}E_{\overline{\nu }_\mathrm{e}}}}{\displaystyle \overline{|(\overline{\nu }_\mathrm{e}+\mathrm{D}\mathrm{e}^++\mathrm{n}+\mathrm{n})|^2}}`$
$`{\displaystyle \frac{1}{2}}(2\pi )^4\delta ^{(4)}(Q+k_{\overline{\nu }_\mathrm{e}}p_1p_2k_{\mathrm{e}^+}){\displaystyle \frac{d^3p_1}{(2\pi )^32E_1}}{\displaystyle \frac{d^3p_2}{(2\pi )^32E_2}}{\displaystyle \frac{d^3k_{\mathrm{e}^+}}{(2\pi )^32E_{\mathrm{e}^+}}},`$ (10.4)
where $`E_\mathrm{D}`$, $`E_{\overline{\nu }_\mathrm{e}}`$, $`E_1`$, $`E_2`$ and $`E_{\mathrm{e}^+}`$ are the energies of the deuteron, the antiโneutrino, the neutrons and the positron. The integration over the phase volume of the ($`\mathrm{nne}^+`$) state we perform in the nonโrelativistic limit and in the rest frame of the deuteron
$`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^3p_1}{(2\pi )^32E_1}\frac{d^3p_2}{(2\pi )^32E_2}\frac{d^3k_{\mathrm{e}^+}}{(2\pi )^32E_{\mathrm{e}^+}}(2\pi )^4\delta ^{(4)}(Q+k_{\overline{\nu }_\mathrm{e}}p_1p_2k_{\mathrm{e}^+})}`$
$`\times {\displaystyle \frac{\left(E_{\mathrm{e}^+}E_{\overline{\nu }_\mathrm{e}}{\displaystyle \frac{1}{3}}\stackrel{}{k}_{\mathrm{e}^+}\stackrel{}{k}_{\overline{\nu }_\mathrm{e}}\right)F_\mathrm{D}^2(M_\mathrm{N}T_{\mathrm{nn}})}{\left(1{\displaystyle \frac{1}{2}}r_{\mathrm{nn}}a_{\mathrm{nn}}M_\mathrm{N}T_{\mathrm{nn}}\right)^2+a_{\mathrm{nn}}^2M_\mathrm{N}T_{\mathrm{nn}}}}={\displaystyle \frac{E_{\overline{\nu }_\mathrm{e}}M_\mathrm{N}^3}{1024\pi ^2}}\left({\displaystyle \frac{E_{\mathrm{th}}}{M_\mathrm{N}}}\right)^{7/2}\left({\displaystyle \frac{2m_\mathrm{e}}{E_{\mathrm{th}}}}\right)^{3/2}{\displaystyle \frac{8}{\pi E_{\mathrm{th}}^2}}`$
$`\times {\displaystyle }{\displaystyle }dT_{\mathrm{e}^+}dT_{\mathrm{nn}}{\displaystyle \frac{\sqrt{T_{\mathrm{e}^+}T_{\mathrm{nn}}}F_\mathrm{D}^2(M_\mathrm{N}T_{\mathrm{nn}})}{\left(1{\displaystyle \frac{1}{2}}r_{\mathrm{nn}}a_{\mathrm{nn}}M_\mathrm{N}T_{\mathrm{nn}}\right)^2+a_{\mathrm{nn}}^2M_\mathrm{N}T_{\mathrm{nn}}}}`$
$`\times (1+{\displaystyle \frac{T_{\mathrm{e}^+}}{m_\mathrm{e}}})\sqrt{1+{\displaystyle \frac{T_{\mathrm{e}^+}}{2m_\mathrm{e}}}}\delta (E_{\overline{\nu }_\mathrm{e}}E_{\mathrm{th}}T_{\mathrm{e}^+}T_{\mathrm{nn}})=`$
$`={\displaystyle \frac{E_{\overline{\nu }_\mathrm{e}}M_\mathrm{N}^3}{1024\pi ^2}}\left({\displaystyle \frac{E_{\mathrm{th}}}{M_\mathrm{N}}}\right)^{7/2}\left({\displaystyle \frac{2m_\mathrm{e}}{E_{\mathrm{th}}}}\right)^{3/2}(y1)^2\mathrm{\Omega }_{\mathrm{nn}\overline{\nu }_\mathrm{e}}(y),`$ (10.5)
where $`T_{\mathrm{nn}}`$ is the kinetic energy of the nn pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state, $`T_{\mathrm{e}^+}`$ and $`m_\mathrm{e}=0.511\mathrm{MeV}`$ are the kinetic energy and the mass of the positron, $`y=E_{\overline{\nu }_\mathrm{e}}/E_{\mathrm{th}}`$ and $`E_{\mathrm{th}}`$ is the antiโneutrino energy threshold of the reaction $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n: $`E_{\mathrm{th}}=\epsilon _\mathrm{D}+m_\mathrm{e}+(M_\mathrm{n}M_\mathrm{p})=(2.225+0.511+1.293)\mathrm{MeV}=4.029\mathrm{MeV}`$. The function $`\mathrm{\Omega }_{\mathrm{nn}\overline{\nu }_\mathrm{e}}(y)`$ is defined by
$`\mathrm{\Omega }_{\mathrm{nn}\overline{\nu }_\mathrm{e}}(y)`$ $`=`$ $`{\displaystyle \frac{8}{\pi }}{\displaystyle \underset{0}{\overset{1}{}}}๐x{\displaystyle \frac{\sqrt{x(1x)}F_\mathrm{D}^2(M_\mathrm{N}E_{\mathrm{th}}(y1)x)}{\left(1{\displaystyle \frac{1}{2}}r_{\mathrm{nn}}a_{\mathrm{nn}}M_\mathrm{N}E_{\mathrm{th}}(y1)x\right)^2+a_{\mathrm{nn}}^2M_\mathrm{N}E_{\mathrm{th}}(y1)x}}`$ (10.6)
$`\times \left(1+{\displaystyle \frac{E_{\mathrm{th}}}{m_\mathrm{e}}}(y1)(1x)\right)\sqrt{1+{\displaystyle \frac{E_{\mathrm{th}}}{2m_\mathrm{e}}}(y1)(1x)},`$
where we have changed the variable $`T_{\mathrm{nn}}=(E_{\overline{\nu }_\mathrm{e}}E_{\mathrm{th}})x`$. The function $`f(y)`$ is normalized to unity at $`y=1`$, i.e. at threshold $`E_{\overline{\nu }_\mathrm{e}}=E_{\mathrm{th}}`$. Thus, the cross section for the reaction $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n reads
$`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(E_{\overline{\nu }_\mathrm{e}})=\sigma _0(y1)^2\mathrm{\Omega }_{\mathrm{nn}\overline{\nu }_\mathrm{e}}(y),`$ (10.7)
where $`\sigma _0`$ amounts to
$`\sigma _0=Q_\mathrm{D}C_{\mathrm{NN}}^2{\displaystyle \frac{9g_\mathrm{A}^2G_\mathrm{V}^2M_\mathrm{N}^8}{512\pi ^4}}\left({\displaystyle \frac{E_{\mathrm{th}}}{M_\mathrm{N}}}\right)^{7/2}\left({\displaystyle \frac{2m_\mathrm{e}}{E_{\mathrm{th}}}}\right)^{3/2}=4.58\times \mathrm{\hspace{0.17em}10}^{43}\mathrm{cm}^2.`$ (10.8)
The experimental data on the antiโneutrino disintegration of the deuteron are given in terms of the cross section averaged over the antiโneutrino energy spectrum :
$`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(E_{\overline{\nu }_\mathrm{e}})_{\mathrm{exp}}=(9.83\pm 2.04)\times 10^{45}\mathrm{cm}^2.`$ (10.9)
In order to average the theoretical cross section Eq.(10.7) over the antiโneutrino spectrum we should use the spectrum given by Table YII of Ref. . This yields
$`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(E_{\overline{\nu }_\mathrm{e}})=11.56\times 10^{45}\mathrm{cm}^2.`$ (10.10)
The theoretical value Eq.(10.10)<sup>4</sup><sup>4</sup>4This result is obtained at zero contribution of the nucleon tensor current (see Appendix and the discussion in the Conclusion). agrees well with the experimental one Eq.(10.9).
## 11 The reactions $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n + p
The amplitude of the neutrino disintegration of the deuteron caused by neutral weak current $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p evaluated through oneโnucleon loop exchanges and at leading order in the large $`N_C`$ expansion reads (see Eq.(A.30) of Appendix):
$`i(\nu _\mathrm{e}+\mathrm{D}\nu _\mathrm{e}+\mathrm{n}+\mathrm{p})=g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}{\displaystyle \frac{G_\mathrm{F}}{\sqrt{2}}}{\displaystyle \frac{3g_\mathrm{V}}{4\pi ^2}}{\displaystyle \frac{1}{1{\displaystyle \frac{1}{2}}r_{\mathrm{np}}a_{\mathrm{np}}k^2+ia_{\mathrm{np}}k}}F_\mathrm{D}(k^2)`$
$`\times e_\mu (k_\mathrm{D})[\overline{u}(k_{\nu _\mathrm{e}}^{})\gamma ^\mu (1\gamma ^5)u(k_{\nu _\mathrm{e}})][\overline{u}(p_2)\gamma ^5u^c(p_1)].`$ (11.1)
The amplitude Eq.(11) squared, averaged over polarizations of the deuteron, summed over polarizations of the nucleons reads
$`\overline{|(\nu _\mathrm{e}+\mathrm{D}\nu _\mathrm{e}+\mathrm{n}+\mathrm{p})|^2}={\displaystyle \frac{36}{\pi ^2}}{\displaystyle \frac{Q_\mathrm{D}g_\mathrm{A}^2G_\mathrm{F}^2C_{\mathrm{NN}}^2M_\mathrm{N}^6F_\mathrm{D}^2(k^2)}{\left(1{\displaystyle \frac{1}{2}}r_{\mathrm{np}}a_{\mathrm{np}}k^2\right)^2+a_{\mathrm{np}}^2k^2}}\left(E_{\nu _\mathrm{e}}^{}E_{\nu _\mathrm{e}}{\displaystyle \frac{1}{3}}\stackrel{}{k}_{\nu _\mathrm{e}}^{}\stackrel{}{k}_{\nu _\mathrm{e}}\right).`$ (11.2)
In the rest frame of the deuteron the cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p is defined by
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(E_{\overline{\nu }_\mathrm{e}})={\displaystyle \frac{1}{4M_\mathrm{D}E_{\nu _\mathrm{e}}}}{\displaystyle \overline{|(\nu _\mathrm{e}+\mathrm{D}\nu _\mathrm{e}+\mathrm{n}+\mathrm{p})|^2}}`$
$`(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\nu _\mathrm{e}}p_1p_2k_{\nu _\mathrm{e}}^{}){\displaystyle \frac{d^3p_1}{(2\pi )^32E_1}}{\displaystyle \frac{d^3p_2}{(2\pi )^32E_2}}{\displaystyle \frac{d^3k_{\nu _\mathrm{e}}^{}}{(2\pi )^32E_{\nu _\mathrm{e}}^{}}}.`$ (11.3)
The integration over the phase volume of the ($`\mathrm{np}\nu _\mathrm{e}`$) state we perform in the nonโrelativistic limit and in the rest frame of the deuteron,
$`{\displaystyle \frac{d^3p_1}{(2\pi )^32E_1}\frac{d^3p_2}{(2\pi )^32E_2}\frac{d^3k_{\nu _\mathrm{e}}^{}}{(2\pi )^32E_{\nu _\mathrm{e}}^{}}(2\pi )^4\delta ^{(4)}(k_\mathrm{D}+k_{\nu _\mathrm{e}}p_1p_2k_{\nu _\mathrm{e}}^{})}`$
$`\left(E_{\nu _\mathrm{e}}E_{\nu _\mathrm{e}}^{}{\displaystyle \frac{1}{3}}\stackrel{}{k}_{\nu _\mathrm{e}}\stackrel{}{k}_{\nu _\mathrm{e}}^{}\right){\displaystyle \frac{F_\mathrm{D}^2(M_\mathrm{N}T_{\mathrm{np}})}{\left(1{\displaystyle \frac{1}{2}}r_{\mathrm{np}}a_{\mathrm{np}}M_\mathrm{N}T_{\mathrm{np}}\right)^2+a_{\mathrm{np}}^2M_\mathrm{N}T_{\mathrm{np}}}}=`$
$`={\displaystyle \frac{E_{\nu _\mathrm{e}}M_\mathrm{N}^3}{210\pi ^3}}\left({\displaystyle \frac{E_{\mathrm{th}}}{M_\mathrm{N}}}\right)^{7/2}(y1)^{7/2}\mathrm{\Omega }_{\mathrm{np}\nu _\mathrm{e}}(y).`$ (11.4)
The function $`\mathrm{\Omega }_{\mathrm{np}\nu _\mathrm{e}}(y)`$, where $`y=E_{\overline{\nu }_\mathrm{e}}/E_{\mathrm{th}}`$ and $`E_{\mathrm{th}}=\epsilon _\mathrm{D}=2.225\mathrm{MeV}`$ is threshold of the reaction, is defined by the integral
$`\mathrm{\Omega }_{\mathrm{np}\nu _\mathrm{e}}(y)={\displaystyle \frac{105}{16}}{\displaystyle \underset{0}{\overset{1}{}}}๐x{\displaystyle \frac{\sqrt{x}(1x)^2}{\left(1{\displaystyle \frac{1}{2}}{\displaystyle \frac{r_{\mathrm{np}}a_{\mathrm{np}}}{r_\mathrm{D}^2}}(y1)x\right)^2+{\displaystyle \frac{a_{\mathrm{np}}^2}{r_\mathrm{D}^2}}(y1)x}}{\displaystyle \frac{1}{(1+(y1)x)^2}},`$ (11.5)
where we have changed the variable $`T_{\mathrm{np}}=(E_{\overline{\nu }_\mathrm{e}}E_{\mathrm{th}})x`$ and used the relation $`M_\mathrm{N}E_{\mathrm{th}}=1/r_\mathrm{D}^2`$ at $`E_{\mathrm{th}}=\epsilon _\mathrm{D}`$. The function $`\mathrm{\Omega }_{\mathrm{np}\nu _\mathrm{e}}(y)`$ is normalized to unity at $`y=1`$, i.e. at threshold $`E_{\overline{\nu }_\mathrm{e}}=E_{\mathrm{th}}`$.
The cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p reads
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(E_{\nu _\mathrm{e}})=\sigma _0(y1)^{7/2}\mathrm{\Omega }_{\mathrm{np}\nu _\mathrm{e}}(y),`$ (11.6)
where $`\sigma _0`$ amounts to
$`\sigma _0=Q_\mathrm{D}C_{\mathrm{NN}}^2{\displaystyle \frac{3g_\mathrm{A}^2G_\mathrm{F}^2M_\mathrm{N}^8}{140\pi ^5}}\left({\displaystyle \frac{E_{\mathrm{th}}}{M_\mathrm{N}}}\right)^{7/2}=1.84\times 10^{43}\mathrm{cm}^2.`$ (11.7)
In our approach the cross section for the disintegration of the deuteron by neutrinos $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p coincides with the cross section for the disintegration of the deuteron by antiโneutrinos $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p, $`\sigma ^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(E_{\nu _\mathrm{e}})=\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\overline{\nu }_\mathrm{e}\mathrm{np}}(E_{\overline{\nu }_\mathrm{e}})`$. Therefore, we have an opportunity to compare our result with the experimental data on the disintegration of the deuteron by antiโneutrinos . The experimental value of the cross section for the antiโneutrino disintegration of the deuteron $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p averaged over the antiโneutrino spectrum reads :
$`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\overline{\nu }_\mathrm{e}\mathrm{np}}(E_{\overline{\nu }_\mathrm{e}})_{\mathrm{exp}}=(6.08\pm 0.77)\times 10^{45}\mathrm{cm}^2.`$ (11.8)
By using the antiโneutrino spectrum given by Table YII of Ref. for the calculation of the average value of the theoretical cross section Eq.(11.6) we obtain
$`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\overline{\nu }_\mathrm{e}\mathrm{np}}(E_{\overline{\nu }_\mathrm{e}})=6.28\times 10^{45}\mathrm{cm}^2.`$ (11.9)
The theoretical value Eq.(11.9)<sup>5</sup><sup>5</sup>5This result is obtained at zero contribution of the nucleon tensor current (see Appendix and the discussion in the Conclusion). agrees well with the experimental one Eq.(11.8).
The cross section for the neutrino disintegration of the deuteron $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p averaged over the $`{}_{}{}^{8}\mathrm{B}`$ solar neutrino spectrum for energy region $`5\mathrm{MeV}E_{\nu _\mathrm{e}}15\mathrm{MeV}`$ is given by
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(E_{\nu _\mathrm{e}})_{\mathrm{\Phi }({}_{}{}^{8}\mathrm{B})}=1.85\times 10^{43}\mathrm{cm}^2.`$ (11.10)
This result can be directly compared with the experimental data obtained by SNO, since the averaged value for the cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p caused by the neutral weak current should not depend on whether neutrino flavours oscillate or not . Of course, if none sterile neutrinos exist .
## 12 Conclusion
We have considered the description of a dynamics of lowโenergy nuclear forces within the NambuโJonaโLasinio model of light nuclei (the NNJL model) for lowโenergy electromagnetic and weak nuclear reactions with the deuteron. We have shown that the NNJL model enables to describe in a reasonable agreement with both experimental data and other theoretical approaches all variety of lowโenergy electromagnetic and weak nuclear reactions with the deuteron in the final or initial state coupled to a nucleonโnucleon (NN) pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state. In the bulk the reaction rates for the neutronโproton radiative capture n + p $``$ D + $`\gamma `$ for thermal neutrons, the photomagnetic disintegration of the deuteron $`\gamma `$ \+ D $``$ n + p, the solar proton burning p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, the pepโprocess p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$, the disintegration of the deuteron by neutrinos and antiโneutrinos caused by charged $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and neutral $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n \+ p weak currents are calculated in agreement with other theorical approaches and experimental data.
When matching our results with those obtained in the PMA we would like to emphasize a much more simple description of the NN interaction responsible for the transition N + N $``$ N + N and a substantial simplification of the evaluation of matrix elements of lowโenergy nuclear transitions near thresholds of the reactions where the NN pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state is produced or absorbed with a relative momentum $`p`$ comparable with zero. Such a simplification is rather clear in the NNJL model where with a good accuracy the deuteron can be considered within a quantum field theoretic approach as a pointโlike particle. Indeed, the spatial region of the localization of the NN pair is of order of $`O(1/p)`$. Near thresholds the effective radius of the deuteron $`r_\mathrm{D}=4.319\mathrm{MeV}`$ is much smaller than $`1/p`$, $`r_\mathrm{D}1/p`$. This yields that the NN pair does not feel the spatial smearing of the deuteron and couples to the deuteron as to a pointโlike particle. A correct description of strong interactions of the pointโlike deuteron coupled to the NN pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state is guaranteed then by the oneโnucleon loop exchanges with dominant contributions of the nucleonโloop anomalies. This implies that the effective Lagrangians $`_{\mathrm{eff}}^{\mathrm{np}\mathrm{D}\gamma }(x)`$, $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$, $`_{\mathrm{eff}}^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(x)`$, $`_{\mathrm{eff}}^{\mathrm{pe}^{}\mathrm{p}\mathrm{D}\nu _\mathrm{e}}(x)`$, $`_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)`$ and $`_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\overline{\nu }_\mathrm{e}\mathrm{np}}(x)`$ are well defined. Thus, the procedure of the derivation of effective Lagrangians of lowโenergy nuclear transitions in the NNJL model treating the deuteron as a pointโlike particle coupled to nucleons and other particles through oneโnucleon loop exchanges seems to be good established. We argue that the application of this procedure should get correct results for the derivation of effective Lagrangians of any other lowโenergy nuclear transitions, effective vertices, of nuclear reactions.
Some problems occur for the evaluation of the amplitudes of nuclear reactions demanding the continuation of matrix elements of lowโenergy nuclear transitions defined by the effective Lagrangians to the energy region far from thresholds.
The continuation of matrix elements of lowโenergy nuclear transitions demands in the NNJL model: 1) the spatial smearing of the NN pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state and 2) the spatial smearing of the deuteron caused by the finite value of the effective radius $`r_\mathrm{D}`$. The spatial smearing of the NN pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state can be carried out by a summation of an infinite series of oneโnucleon bubbles describing rescattering or differently a relative movement of the NN pair either in an initial or a final state of a nuclear reaction. The result of the NN rescattering can be expressed in terms of Sโwave scattering lengths and effective ranges in complete agreement with nuclear phenomenology of lowโenergy elastic NN scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state. However, for the description of the spatial smearing of the deuteron the abilities of the NNJL model are rescticted and most what one can do at present level of the development of the model is to introduce the spatial smearing of the deuteron phenomenologically in the form of the Fourier transform of the approximate $`{}_{}{}^{3}\mathrm{S}_{1}^{}`$ wave function of the deuteron normalized to unity at $`p=0`$: $`F_\mathrm{D}(p^2)=1/(1+r_\mathrm{D}^2p^2)`$. We have chosen a simplest form of the spatial smearing of the deuteron. Of course, $`F_\mathrm{D}(p^2)`$ can be taken in the more complicated form of the Fourier transform of the explicit wave function of the deuteron. Of course, such a dependence is not absolute and the spatial smearing of the deuteron can be taken into account in the form of phenomenological form factors as it has been done by Mintz , for example.
We would like to emphasize that the cross section for the M1โcapture n + p $``$ D + $`\gamma `$ for thermal neutrons has been calculated by accounting for the contributions of chiral oneโmeson loop corrections and the $`\mathrm{\Delta }(1232)`$ resonance. The total cross section for the M1โcapture has been found dependent on the parameter $`Z`$ defining the $`\pi \mathrm{N}\mathrm{\Delta }`$ coupling constant offโmass shell of the $`\mathrm{\Delta }(1232)`$ resonance: $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=325.5(1+0.246(12Z))^2\mathrm{mb}`$ Eq.(4.15). In order to fit the experimental value of the cross section we should set $`Z=0.473`$. This agrees with the experimental bound $`|Z|1/2`$ . At $`Z=1/2`$ the contribution of the $`\mathrm{\Delta }(1232)`$ resonance to the amplitude of the M1โcapture is defined only by the nucleon tensor current. Setting $`Z=1/2`$ that is favoured theoretically we calculate the cross section for the M1โcapture $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=325.5\mathrm{mb}`$ agreeing with the experimental data with accuracy better than 3$`\%`$.
When matching our result for the cross section for the M1โcapture with the recent one obtained in the EFT approach by Chen, Rupak and Savage : $`\sigma (\mathrm{np}\mathrm{D}\gamma )(T_\mathrm{n})=(287.1+6.51{}_{}{}^{\mathit{\pi ฬธ}}L_{1}^{})\mathrm{mb}`$ (see Eq.(3.49) of Ref. ), we accentuate the dependence of the cross section on the parameter $`{}_{}{}^{\mathit{\pi ฬธ}}L_{1}^{}`$ undefined in the approach. This parameter has to be fixed from the experimental data . In the NNJL model the parameter $`{}_{}{}^{\mathit{\pi ฬธ}}L_{1}^{}`$ can be expressed in terms of the $`Z`$ parameter as follows: $`{}_{}{}^{\mathit{\pi ฬธ}}L_{1}^{}=5.90+24.60(12Z)+3.03(12Z)^2`$. Thus, in the NNJL model the parameter $`{}_{}{}^{\mathit{\pi ฬธ}}L_{1}^{}`$ acquires a distinct meaning of the contribution of the $`\mathrm{\Delta }(1232)`$ resonance.
The obtained result for the M1โcapture n + p $``$ D + $`\gamma `$ we have applied to the analysis of the photomagnetic disintegration of the deuteron $`\gamma `$ \+ D $``$ n + p. Due to timeโreversal invariance the cross section for the photomagnetic disintegration of the deuteron can be directly expressed through the cross section for the M1โcapture. We have compared the numerical values of the cross section $`\sigma (\gamma \mathrm{D}\mathrm{np})(\omega )`$ calculated in the NNJL model with the results obtained by Chen and Savage and found a good agreement. Nevertheless, it should be emphasized that in the critical region of the photon energies $`\omega 2\epsilon _\mathrm{D}=4.45\mathrm{MeV}`$ restricting the energy region of the dominance of the photomagnetic disintegration of the deuteron the cross section calculated in the NNJL model falls steeper than the cross section obtained in the EFT. However, in this region the dominant role should be attributed to the E1โtransition , which we have not considered. The comparison of our results on the photomagnetic disintegration of the deuteron with the experimental data demands the inclusion of the E1โtransition as well. This is planned in our forthcoming publications.
The effective coupling constants of lowโenergy weak transitions p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$, $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p, $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n + p have been found dependent on an arbitrary parameter $`\overline{\xi }`$ in the form $`g_\mathrm{V}g_\mathrm{V}(1+\overline{\xi })`$ caused by the contribution of the nucleon tensor current .
Since at $`\overline{\xi }=0`$ we get the value of the astrophysical factor $`S_{\mathrm{pp}}(0)=4.08\times 10^{25}\mathrm{MeV}\mathrm{b}`$ in complete agreement with the recommended one $`S_{\mathrm{pp}}(0)=4.00\times 10^{25}\mathrm{MeV}\mathrm{b}`$ related to the Standard Solar Model , any nonโzero value of $`\overline{\xi }`$ should lead to an Alternative Solar Model (ASM). The problem of the formulation of the ASM goes beyond the scope of this paper. Therefore, in order to dwell within the Standard Solar Model we have to set $`\overline{\xi }=0`$.
Setting $`\overline{\xi }=0`$ we have shown that all lowโenergy weak nuclear reactions of astrophysical interest (i) the solar proton burning p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, (ii) the pepโprocess p \+ e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$ and (iii) the disintegration of the deuteron by neutrinos and antiโneutrinos caused by charged $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and neutral $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n + p weak currents are described in the bulk in agreement with other theorical approaches and experimental data. The effective Lagrangians of lowโenergy weak nuclear transitions p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$, $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p, $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n \+ p are determined by the anomaly of the oneโnucleon loop triangle $`AAV`$โdiagrams. This confirms the statement argued in the NNJL model concerning the dominant role of the oneโnucleon loop anomalies for the description of lowโenergy nuclear forces within a quantum field theoretic approach.
We have shown that the contributions of lowโenergy elastic pp scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state with the Coulomb repulsion to the amplitudes of the reactions p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p and p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$ are described in the NNJL model in full agreement with lowโenergy nuclear phenomenology in terms of the Sโwave scattering length and the effective range. The amplitude of lowโenergy elastic pp scattering has been obtained by summing up an infinite series of oneโproton loop diagrams and evaluating the result of the summation at leading order in the large $`N_C`$ expansion. The same method has been applied to the evaluation of the contribution of lowโenergy nuclear forces to the relative movements of the nn and np pairs, respectively, for the reactions $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ D $``$ $`\nu _\mathrm{e}(\overline{\nu }_\mathrm{e})`$ \+ n + p. This has given the amplitudes of lowโenergy elastic nn and np scattering described in terms of Sโwave scattering lengths and effective ranges in agreement with lowโenergy nuclear phenomenology as well.
The astrophysical factor $`S_{\mathrm{pep}}(0)`$ for the pepโprocess, p + e<sup>-</sup> \+ p $``$ D + $`\nu _\mathrm{e}`$, evaluated relative to $`S_{\mathrm{pp}}(0)`$ is found in full agreement with the result obtained by Bahcall and May .
The cross sections for the antiโneutrino disintegration of the deuteron caused by charged $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and neutral $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p weak currents and averaged over the antiโneutrino spectrum $`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(E_{\overline{\nu }_\mathrm{e}})=11.56\times 10^{45}\mathrm{cm}^2`$ and $`\sigma ^{\overline{\nu }_\mathrm{e}\mathrm{D}\overline{\nu }_\mathrm{e}\mathrm{np}}(E_{\overline{\nu }_\mathrm{e}})=6.28\times 10^{45}\mathrm{cm}^2`$ agree well with recent experimental data
$`\sigma ^{\overline{\nu }_\mathrm{e}De^+nn}(E_{\overline{\nu }_\mathrm{e}})_{\mathrm{exp}}`$ $`=`$ $`(9.83\pm 2.04)\times 10^{45}\mathrm{cm}^2,`$
$`\sigma ^{\overline{\nu }_\mathrm{e}D\overline{\nu }_\mathrm{e}np}(E_{\overline{\nu }_\mathrm{e}})_{\mathrm{exp}}`$ $`=`$ $`(6.08\pm 0.77)\times 10^{45}\mathrm{cm}^2`$
obtained by the Reinesโs experimental group .
The cross sections for the reactions $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p have been recently calculated by Butler and Chen in the EFT approach. The obtained results have been written in the following general form $`\sigma =(a+bL_{1,\mathrm{A}})\times 10^{42}\mathrm{cm}^2`$ (see Table I of Ref. ), where $`a`$ and $`b`$ are the parameters which have been calculated in the approach, whereas $`L_{1,\mathrm{A}}`$ is a free one. In the NNJL model the appearance of the free parameter is related to the contribution of the nucleon tensor current that effectively leads to the change of the coupling constant $`g_\mathrm{V}g_\mathrm{V}(1+\overline{\xi })`$, where $`\overline{\xi }`$ is an arbitrary parameter. The best agreement with the recommended value of the astrophysical factor for the solar proton burning and the contemporary experimental data on the cross sections for the antiโneutrino disintegration of the deuteron $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p caused by charged and neutral weak current, respectively, we obtain at $`\overline{\xi }=0`$ (see Appendix).
We would like to emphasize that the contribution of the $`\mathrm{\Delta }(1232)`$ resonance to the amplitudes of the lowโenergy weak nuclear reactions p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, p + e<sup>-</sup> \+ p $``$ $`\nu _\mathrm{e}`$ \+ D, $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p, $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p can be neglected. In fact, the contribution of the $`\mathrm{\Delta }(1232)`$ resonance to the amplitudes of these reactions is of order of the momentum of the leptonic pair. This is due to the gauge invariance of the effective interactions $`\mathrm{\Delta }NW^+`$ and $`\mathrm{\Delta }NZ`$ which should be proportional to $`W_{\mu \nu }^+=_\mu W_\nu ^+_\nu W_\mu ^+`$ and $`Z_{\mu \nu }=_\mu Z_\nu _\nu Z_\mu `$, respectively. Since the amplitudes of the lowโenergy weak nuclear reactions under consideration are defined by the GamowโTeller transitions, the terms proportional to the momentum of the leptonic pair give negligible small contributions. Thus, the cross sections for lowโenergy weak nuclear reactions enumerated above do not depend practically on the uncertainties of the parameter $`Z`$.
This distinguishes lowโenergy weak nuclear reactions with the deuteron from the neutronโproton radiative capture analysed in the NNJL model. Unlike the lowโenergy weak nuclear reactions the contribution of the $`\mathrm{\Delta }(1232)`$ resonance to the amplitude of the neutronโproton radiative capture is essential for the explanation of the experimental data.
The cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p has been evaluated with respect to $`S_{\mathrm{pp}}(0)`$. We have found an enhancement of the cross section by a factor of 1.6 on the average relative to the results obtained in the PMA. It would be important to verify this result for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p \+ p in solar neutrino experiments planned by SNO. Indeed, first, this should provide an experimental study of $`S_{\mathrm{pp}}(0)`$ and, second, the cross sections for the antiโneutrino disintegration of the deuteron caused by charged $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and neutral $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p weak currents have been found in good agreement with recent experimental data .
For the comparison of the cross sections for the reactions $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p and $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p caused by charged and neutral weak currents, respectively, we have averaged the cross sections over the $`{}_{}{}^{8}\mathrm{B}`$ solar neutrino energy spectrum at energy region $`5\mathrm{MeV}E_{\nu _\mathrm{e}}15\mathrm{MeV}`$, where the lower bound is related to the experimental threshold of the experiments at SNO and the upper one is defined by the kinematics of the $`\beta `$ decay $`{}_{}{}^{8}\mathrm{B}{}_{}{}^{8}\mathrm{Be}_{}^{}`$ \+ e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ being the source of the $`{}_{}{}^{8}\mathrm{B}`$ neutrinos in the solar core. We have obtained
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\mathrm{e}^{}\mathrm{pp}}(E_{\nu _\mathrm{e}})_{\mathrm{\Phi }({}_{}{}^{8}\mathrm{B})}`$ $`=`$ $`2.62\times 10^{42}\mathrm{cm}^2,`$
$`\sigma ^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(E_{\nu _\mathrm{e}})_{\mathrm{\Phi }({}_{}{}^{8}\mathrm{B})}`$ $`=`$ $`1.85\times 10^{43}\mathrm{cm}^2.`$
The experimental value for the cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p caused by the charged weak current can, in principle, differ from the theoretical one due to a possible contribution of the neutrino flavour oscillations . In turn, the averaged value of the cross section for the reaction $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p caused by the neutral weak current should be directly compared with the experimental data, since it should not depend on whether neutrino flavours oscillate or not. Of course, the former is valid only if there is no soโcalled sterile neutrino having no interactions with Standard Model particles .
Concluding the paper we would like to emphasize that the NNJL model conveys the idea of a dominant role of oneโfermion loop (oneโnucleon loop) anomalies from elementary particle physics to the nuclear one. This is a new approach to the description of lowโenergy nuclear forces in physics of light nuclei. In spite of almost 30 years have passed after the discovery of oneโfermion loop anomalies and application of these anomalies to the evaluation of effective chiral Lagrangians of lowโenergy interactions of hadrons, in nuclear physics fermionโloop anomalies have not been applied to the analysis of lowโenergy nuclear interactions and properties of light nuclei. The contributions of oneโnucleon loop anomalies are strongly related to highโenergy $`N\overline{N}`$ fluctuations of virtual nucleon fields . An important role of $`N\overline{N}`$ fluctuations for the correct description of lowโenergy properties of finite nuclei has been understood in Ref. . Moreover, $`N\overline{N}`$ fluctuations have been described in terms of oneโnucleon loop diagrams within quantum field theoretic approaches, but the contributions of oneโnucleon loop anomalies have not been considered. The NNJL model allows to fill this blank. Within the framework of the NNJL model we aim to understand, in principle, the role of nucleonโloop anomalies for the description of a dynamics of low-energy nuclear forces at the quantum field theoretic level.
## Acknowledgement
We are grateful to Prof. M. Kamionkowski for helpful remarks and encouragement for further applications of the expounded in the paper technique to the evaluation of the astrophysical factor for pp fusion and the cross section for the neutrino disintegration of the deuteron $`\nu _\mathrm{e}`$ \+ D $``$ e<sup>-</sup> \+ p + p by accounting for the Coulomb repulsion between the protons. We thank Prof. J. N. Bahcall for discussions concerning the $`{}_{}{}^{8}\mathrm{B}`$ neutrino energy spectrum and experiments at SNO. We thank Dr. V. A. Sadovnikova for many helpful and interesting discussions.
We thank Dr. J. Beacom for calling our attention to the experimental data . Discussions of the experimental data with Prof. H. Sobel and Dr. L. Price are greatly appreciated.
## Appendix. The effective Lagrangian of the transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$
The reaction p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ runs through the intermediate Wโboson exchange, p + p $``$ D + W<sup>+</sup> $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$. In the NNJL model we determine this transition in terms of the following effective interactions
$$_{\mathrm{npD}}(x)=ig_\mathrm{V}[\overline{p^c}(x)\gamma ^\mu n(x)\overline{n^c}(x)\gamma ^\mu p(x)]D_\mu ^{}(x),$$
$$_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x)=\frac{1}{2}C_{\mathrm{NN}}\{[\overline{p}(x)\gamma ^\mu \gamma ^5p^c(x)][\overline{p^c}(x)\gamma _\mu \gamma ^5p(x)]+[\overline{p}(x)\gamma ^5p^c(x)][\overline{p^c}(x)\gamma ^5p(x)]\},$$
$$_{\mathrm{npW}}(x)=\frac{g_\mathrm{W}}{2\sqrt{2}}\mathrm{cos}\vartheta _C[\overline{n}(x)\gamma ^\nu (1g_\mathrm{A}\gamma ^5)p(x)]W_\nu ^{}(x).$$
$`(\mathrm{A}.1)`$
The transition W<sup>+</sup> $``$ e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ is defined by the Lagrangian
$$_{\nu _\mathrm{e}\mathrm{e}^+\mathrm{W}}(x)=\frac{g_\mathrm{W}}{2\sqrt{2}}[\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)]W_\nu ^+(x).$$
$`(\mathrm{A}.2)`$
The electroweak coupling constant $`g_\mathrm{W}`$ is connected with the Fermi weak constant $`G_\mathrm{F}`$ and the mass of the W boson $`M_\mathrm{W}`$ through the relation
$$\frac{g_\mathrm{W}^2}{8M_\mathrm{W}^2}=\frac{G_\mathrm{F}}{\sqrt{2}}.$$
$`(\mathrm{A}.3)`$
For the evaluation of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ it is convenient to use the interaction
$$_{\mathrm{npW}}(x)=[\overline{n}(x)\gamma ^\nu (1g_\mathrm{A}\gamma ^5)p(x)]W_\nu ^{}(x),$$
$`(\mathrm{A}.4)`$
and for the description of the subsequent weak transition W<sup>+</sup> $``$ e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ to replace the operator of the Wโboson field by the operator of the leptonic weak current
$$W_\nu ^{}(x)\frac{G_\mathrm{V}}{\sqrt{2}}[\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma _\nu (1\gamma ^5)\psi _\mathrm{e}(x)],$$
$`(\mathrm{A}.5)`$
where $`G_\mathrm{V}=G_\mathrm{F}\mathrm{cos}\vartheta _C`$.
The S matrix describing the transitions like p + p $``$ D + W<sup>+</sup> is defined by
$$\mathrm{S}=\mathrm{T}e^{i{\scriptscriptstyle d^4x\left[_{\mathrm{npD}}\left(x\right)+_{\mathrm{npW}}\left(x\right)+_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}\left(x\right)+\mathrm{}\right]}},$$
$`(\mathrm{A}.6)`$
where T is the timeโordering operator and the ellipses denote the contribution of interactions irrelevant for the problem.
For the evaluation of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)`$ we have to consider the third order term of the S matrix which reads
$$\mathrm{S}^{(3)}=\frac{i^3}{3!}d^4x_1d^4x_2d^4x_3\mathrm{T}([_{\mathrm{npD}}(x_1)+_{\mathrm{npW}}(x_1)+_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)+\mathrm{}]$$
$$\times [_{\mathrm{npD}}(x_2)+_{\mathrm{npW}}(x_2)+_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_2)+\mathrm{}]$$
$$\times [_{\mathrm{npD}}(x_3)+_{\mathrm{npW}}(x_3)+_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_3)+\mathrm{}])=$$
$$=id^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)_{\mathrm{npD}}(x_2)_{\mathrm{npW}}(x_3))+\mathrm{}$$
$`(\mathrm{A}.7)`$
The ellipses denote the terms which do not contribute to the transition p + p $``$ D + W<sup>+</sup> and the interaction $`_{\mathrm{npW}}(x)`$. The S matrix element $`\mathrm{S}_{\mathrm{pp}\mathrm{DW}^+}^{(3)}`$ describing the transition p + p $``$ D + W<sup>+</sup> we determine as follows
$$\mathrm{S}_{\mathrm{pp}\mathrm{DW}^+}^{(3)}=id^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)_{\mathrm{npD}}(x_2)_{\mathrm{npW}}(x_3)).$$
$`(\mathrm{A}.8)`$
For the derivation of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)`$ from the S matrix element Eq.(A.8) we should make all necessary contractions of the operators of the proton and the neutron fields. These contractions we denote by the brackets as
$$\mathrm{S}_{\mathrm{pp}\mathrm{DW}^+}^{(3)}=id^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)_{\mathrm{npD}}(x_2)_{\mathrm{npW}}(x_3)).$$
$`(\mathrm{A}.9)`$
Now the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)`$ related to the S matrix element $`\mathrm{S}_{\mathrm{pp}\mathrm{DW}^+}^{(3)}`$ can be defined by
$$\mathrm{S}_{\mathrm{pp}\mathrm{DW}^+}^{(3)}=id^4x_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)=$$
$$=id^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)_{\mathrm{npD}}(x_2)_{\mathrm{npW}}(x_3)).$$
$`(\mathrm{A}.10)`$
In terms of the operators of the interacting fields the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)`$ reads
$$d^4x_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)=d^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)_{\mathrm{npD}}(x_2)_{\mathrm{npW}}(x_3))$$
$$=\frac{1}{2}C_{\mathrm{NN}}\times (ig_\mathrm{V})\times (g_\mathrm{A})d^4x_1d^4x_2d^4x_3\mathrm{T}([\overline{p^c}(x_1)\gamma _\alpha \gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))$$
$$\times 0|\mathrm{T}([\overline{p}(x_1)\gamma ^\alpha \gamma ^5p^c(x_1)][\overline{p^c}(x_2)\gamma ^\mu n(x_2)\overline{n^c}(x_2)\gamma ^\mu p(x_2)][\overline{n}(x_3)\gamma ^\nu \gamma ^5p(x_3)])|0$$
$$\frac{1}{2}C_{\mathrm{NN}}\times (ig_\mathrm{V})\times (g_\mathrm{A})d^4x_1d^4x_2d^4x_3\mathrm{T}([\overline{p^c}(x_1)\gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))$$
$$\times 0|\mathrm{T}([\overline{p}(x_1)\gamma ^5p^c(x_1)][\overline{p^c}(x_2)\gamma ^\mu n(x_2)\overline{n^c}(x_2)\gamma ^\mu p(x_2)][\overline{n}(x_3)\gamma ^\nu \gamma ^5p(x_3)])|0.$$
$`(\mathrm{A}.11)`$
Since p + p $``$ D + W<sup>+</sup> is the GamowโTeller transition, we have taken into account the Wโboson coupled to the axialโvector nucleon current.
Due to the relation $`\overline{n^c}(x_2)\gamma ^\mu p(x_2)=\overline{p^c}(x_2)\gamma ^\mu n(x_2)`$ the r.h.s. of Eq.(A.11) can be reduced as follows
$$d^4x_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)=d^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)_{\mathrm{npD}}(x_2)_{\mathrm{npW}}(x_3))$$
$$=C_{\mathrm{NN}}\times (ig_\mathrm{V})\times g_\mathrm{A}d^4x_1d^4x_2d^4x_3\mathrm{T}([\overline{p^c}(x_1)\gamma _\alpha \gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))$$
$$\times 0|\mathrm{T}([\overline{p}(x_1)\gamma ^\alpha \gamma ^5p^c(x_1)][\overline{p^c}(x_2)\gamma ^\mu n(x_2)][\overline{n}(x_3)\gamma ^\nu \gamma ^5p(x_3)])|0$$
$$+C_{\mathrm{NN}}\times (ig_\mathrm{V})\times g_\mathrm{A}d^4x_1d^4x_2d^4x_3\mathrm{T}([\overline{p^c}(x_1)\gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))$$
$$\times 0|\mathrm{T}([\overline{p}(x_1)\gamma ^5p^c(x_1)][\overline{p^c}(x_2)\gamma ^\mu n(x_2)][\overline{n}(x_3)\gamma ^\nu \gamma ^5p(x_3)])|0.$$
$`(\mathrm{A}.12)`$
Making the necessary contractions we arrive at the expression
$$d^4x_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)=d^4x_1d^4x_2d^4x_3\mathrm{T}(_{\mathrm{eff}}^{\mathrm{pp}\mathrm{pp}}(x_1)_{\mathrm{npD}}(x_2)_{\mathrm{npW}}(x_3))=$$
$$=2\times C_{\mathrm{NN}}\times (ig_\mathrm{V})\times g_\mathrm{A}d^4x_1d^4x_2d^4x_3\mathrm{T}([\overline{p^c}(x_1)\gamma _\alpha \gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))$$
$$\times (1)\mathrm{tr}\{\gamma ^\alpha \gamma ^5(i)S_F^c(x_1x_2)\gamma ^\mu (i)S_F(x_2x_3)\gamma ^\nu \gamma ^5(i)S_F(x_3x_1)\}$$
$$+2\times C_{\mathrm{NN}}\times (ig_\mathrm{V})\times g_\mathrm{A}d^4x_1d^4x_2d^4x_3\mathrm{T}([\overline{p^c}(x_1)\gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))$$
$$\times (1)\mathrm{tr}\{\gamma ^5(i)S_F^c(x_1x_2)\gamma ^\mu (i)S_F(x_2x_3)\gamma ^\nu \gamma ^5(i)S_F(x_3x_1)\},$$
$`(\mathrm{A}.13)`$
where the combinatorial factor 2 takes into account the fact that the protons are identical particles in the nucleon loop.
In the momentum representation of the Green functions we get
$$d^4x_{\mathrm{eff}}^{\mathrm{pp}\mathrm{DW}^+}(x)=$$
$$=ig_\mathrm{A}C_{\mathrm{NN}}\frac{g_\mathrm{V}}{8\pi ^2}d^4x_1\frac{d^4x_2d^4k_2}{(2\pi )^4}\frac{d^4x_3d^4k_3}{(2\pi )^4}e^{ik_2\left(x_2x_1\right)}e^{ik_3\left(x_3x_1\right)}$$
$$\times \mathrm{T}([\overline{p^c}(x_1)\gamma _\alpha \gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))J^{\alpha \mu \nu }(k_2,k_3;Q)$$
$$ig_\mathrm{A}C_{\mathrm{NN}}\frac{g_\mathrm{V}}{8\pi ^2}d^4x_1\frac{d^4x_2d^4k_2}{(2\pi )^4}\frac{d^4x_3d^4k_3}{(2\pi )^4}e^{ik_2\left(x_2x_1\right)}e^{ik_3\left(x_3x_1\right)}$$
$$\times \mathrm{T}([\overline{p^c}(x_1)\gamma ^5p(x_1)]D_\mu ^{}(x_2)W_\nu ^{}(x_3))J^{\mu \nu }(k_2,k_3;Q).$$
$`(\mathrm{A}.14)`$
The structure functions $`๐ฅ^{\alpha \mu \nu }(k_2,k_3;Q)`$ and $`๐ฅ^{\mu \nu }(k_2,k_3;Q)`$ are defined by the momentum integrals
$$๐ฅ^{\alpha \mu \nu }(k_2,k_3;Q)=$$
$$=\frac{d^4k}{\pi ^2i}\mathrm{tr}\left\{\gamma ^\alpha \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}+\widehat{k}_2}\gamma ^\mu \frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\gamma ^\nu \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}\widehat{k}_3}\right\},$$
$$๐ฅ^{\mu \nu }(k_2,k_3;Q)=$$
$$=\frac{d^4k}{\pi ^2i}\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}+\widehat{k}_2}\gamma ^\mu \frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\gamma ^\nu \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}\widehat{k}_3}\right\}.$$
$`(\mathrm{A}.15)`$
We have introduced a 4โvector $`Q=ak_2+bk_3`$ caused by an arbitrary shift of a virtual momentum with arbitrary parameters $`a`$ and $`b`$.
In order to obtain the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ of the transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ we have to replace the operator of the Wโboson field by the operator of the leptonic weak current Eq.(A.5):
$$d^4x_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=$$
$$=ig_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{g_\mathrm{V}}{8\pi ^2}d^4x_1\frac{d^4x_2d^4k_2}{(2\pi )^4}\frac{d^4x_3d^4k_3}{(2\pi )^4}e^{ik_2\left(x_2x_1\right)}e^{ik_3\left(x_3x_1\right)}$$
$$\times \mathrm{T}([\overline{p^c}(x_1)\gamma _\alpha \gamma ^5p(x_1)]D_\mu ^{}(x_2)[\overline{\psi }_{\nu _\mathrm{e}}(x_3)\gamma _\nu (1\gamma ^5)\psi _\mathrm{e}(x_3)])J^{\alpha \mu \nu }(k_2,k_3;Q)$$
$$+ig_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{g_\mathrm{V}}{8\pi ^2}d^4x_1\frac{d^4x_2d^4k_2}{(2\pi )^4}\frac{d^4x_3d^4k_3}{(2\pi )^4}e^{ik_2\left(x_2x_1\right)}e^{ik_3\left(x_3x_1\right)}$$
$$\times \mathrm{T}([\overline{p^c}(x_1)\gamma ^5p(x_1)]D_\mu ^{}(x_2)[\overline{\psi }_{\nu _\mathrm{e}}(x_3)\gamma _\nu (1\gamma ^5)\psi _\mathrm{e}(x_3)])J^{\mu \nu }(k_2,k_3;Q).$$
$`(\mathrm{A}.16)`$
Thus, the problem of the evaluation of the effective Lagrangian of the transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ reduces itself to the problem of the evaluation of the structure functions Eq.(A.15). The momentum $`k_3`$ is related to the 4โmomentum of the leptonic pair. Due to the GamowโTeller type of the transition p + p $``$ D + W<sup>+</sup> the contribution proportional to the 4โmomentum of the leptonic pair turns out to be much smaller with respect to the contribution proportional to the 4โmomentum of the deuteron $`k_2`$. Therefore, without loss of generality we can set $`k_3=0`$ in the integrand. This gives
$$๐ฅ^{\alpha \mu \nu }(k_2,k_3;Q)=$$
$$=\frac{d^4k}{\pi ^2i}\mathrm{tr}\left\{\gamma ^\alpha \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}+\widehat{k}_2}\gamma ^\mu \frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\gamma ^\nu \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\right\},$$
$$๐ฅ^{\mu \nu }(k_2,k_3;Q)=$$
$$=\frac{d^4k}{\pi ^2i}\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}+\widehat{k}_2}\gamma ^\mu \frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\gamma ^\nu \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\right\},$$
$`(\mathrm{A}.17)`$
The evaluation of the momentum integrals at leading order in the $`1/M_\mathrm{N}`$ expansion corresponding the leading order in the large $`N_C`$ expansion due to the proportionality $`M_\mathrm{N}N_C`$ in QCD with $`SU(N_C)`$ gauge group at $`N_C\mathrm{}`$ (see also Ref. ) yields
$$๐ฅ^{\alpha \mu \nu }(k_2,k_3;Q)=3(k_2^\alpha g^{\nu \mu }k_2^\nu g^{\mu \alpha })+\frac{1}{9}(1+2a)(k_2^\alpha g^{\nu \mu }+k_2^\nu g^{\mu \alpha }),$$
$$๐ฅ^{\mu \nu }(k_2,k_3;Q)=g^{\mu \nu }4M_\mathrm{N}J_2(M_\mathrm{N})O(1/N_C^2),$$
$`(\mathrm{A}.18)`$
where the terms proportional to $`k_2^\mu `$ have been dropped, because they produce the contributions to the effective Lagrangian multiplied by $`^\mu D_\mu (x)`$ vanishing by virtue of the constraint $`^\mu D_\mu (x)=0`$. Then, $`J_2(M_\mathrm{N})`$ is a logarithmically divergent integral defined in the NNJL model in terms of the cutโoff $`\mathrm{\Lambda }_D=46.172\mathrm{MeV}`$ such as $`\mathrm{\Lambda }_\mathrm{D}M_\mathrm{N}`$ :
$$J_2(M_\mathrm{N})=\frac{d^4k}{\pi ^2i}\frac{1}{(M_\mathrm{N}^2k^2)^2}=2\underset{0}{\overset{\mathrm{\Lambda }_\mathrm{D}}{}}\frac{d|\stackrel{}{k}|\stackrel{}{k}^{\mathrm{\hspace{0.17em}2}}}{(M_\mathrm{N}^2+\stackrel{}{k}^{\mathrm{\hspace{0.17em}2}})^{3/2}}=\frac{2}{3}\left(\frac{\mathrm{\Lambda }_\mathrm{D}}{M_\mathrm{N}}\right)^3O(1/N_C^3).$$
$`(\mathrm{A}.19)`$
The cutโoff $`\mathrm{\Lambda }_\mathrm{D}`$ restricts 3โmomenta of the virtual nucleon fluctuations forming the physical deuteron . Due to the uncertainty relation $`\mathrm{\Delta }r\mathrm{\Lambda }_\mathrm{D}1`$ the spatial region of virtual nucleon fluctuations forming the physical deuteron is defined by $`\mathrm{\Delta }r4.274\mathrm{fm}`$. This agrees with the effective radius of the deuteron $`r_\mathrm{D}=1/\sqrt{\epsilon _\mathrm{D}M_\mathrm{N}}=4.319\mathrm{fm}`$.
Keeping the terms of the same order in the large $`N_C`$ expansion we get the structure functions
$$๐ฅ^{\alpha \mu \nu }(k_2,k_3;Q)=3(k_2^\alpha g^{\nu \mu }k_2^\nu g^{\mu \alpha })+\frac{1}{9}(1+2a)(k_2^\alpha g^{\nu \mu }+k_2^\nu g^{\mu \alpha }),$$
$$๐ฅ^{\mu \nu }(k_2,k_3;Q)=0,$$
$`(\mathrm{A}.20)`$
The structure functions Eq.(A.20) define the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$:
$$_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=g_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{V}}{8\pi ^2}[\overline{p^c}(x)\gamma ^\mu \gamma ^5p(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)]$$
$$\times \left[(_\mu D_\nu ^{}(x)_\nu D_\mu ^{}(x))+\frac{1}{27}(1+2a)(_\mu D_\nu ^{}(x)+_\nu D_\mu ^{}(x))\right].$$
$`(\mathrm{A}.21)`$
In order to remove uncertainty caused by the shift of virtual momenta of the oneโnucleon loop diagrams we demand the invariance of the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ under gauge transformations of the deuteron field
$$D_\mu ^{}(x)D_\mu ^{}(x)+_\mu \mathrm{\Omega }(x),$$
$`(\mathrm{A}.22)`$
where $`\mathrm{\Omega }(x)`$ is a gauge function. The requirement of the invariance of the effective Lagrangian of the lowโenergy transition p \+ p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ under gauge transformation Eq.(A.22) imposes the constraint $`a=1/2`$. This reduces the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ to the form
$$_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=g_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{V}}{8\pi ^2}D_{\mu \nu }^{}(x)[\overline{p^c}(x)\gamma ^\mu \gamma ^5p(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)].$$
$`(\mathrm{A}.23)`$
This effective Lagrangian is defined by the structure function $`๐ฅ^{\alpha \mu \nu }(k_2,k_2;Q)`$:
$$๐ฅ^{\alpha \mu \nu }(k_2,k_3;Q)=3(k_2^\alpha g^{\nu \mu }k_2^\nu g^{\mu \alpha }).$$
$`(\mathrm{A}.24)`$
The structure function $`๐ฅ^{\alpha \mu \nu }(k_2,k_2;Q)`$ does not depend on the mass of virtual nucleons and according to Gertsein and Jackiw can be valued as the anomaly of the oneโnucleon triangle $`AAV`$โdiagram. The requirement of gauge invariance applied to remove ambiguities of the structure function $`๐ฅ^{\alpha \mu \nu }(k_2,k_3;Q)`$ and to fix the contribution of the anomaly of the oneโnucleon loop $`AAV`$โdiagrams is in complete agreement with the derivation of the AdlerโBellโJackiw axialโvector anomaly performed in terms of oneโnucleon loop $`AVV`$โdiagrams (see Jackiw ).
The effective Lagrangian $`_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)`$ describing the lowโenergy transition $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n can be obtained by the way analogous to $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ and reads
$$_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)=g_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{V}}{8\pi ^2}D_{\mu \nu }(x)[\overline{n}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)].$$
$`(\mathrm{A}.25)`$
In the lowโenergy limit when $`D_{\mu \nu }^{}(x)[\overline{p^c}(x)\gamma ^\mu \gamma ^5p(x)]\mathrm{\hspace{0.17em}2}iM_\mathrm{N}D_\nu ^{}(x)[\overline{p^c}(x)\gamma ^5p(x)]`$ the effective Lagrangian Eq.(A.23) reduces itself to the form
$$_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=ig_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{V}}{4\pi ^2}D_\nu ^{}(x)[\overline{p^c}(x)\gamma ^5p(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)].$$
$`(\mathrm{A}.26)`$
The lowโenergy reduction $`D_{\mu \nu }[\overline{n}(x)\gamma ^\mu \gamma ^5n^c(x)]\mathrm{\hspace{0.17em}2}iM_\mathrm{N}D_\nu (x)\overline{n}(x)\gamma ^5n^c(x)`$ applied to the effective Lagrangian $`_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)`$ gives
$$_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)=ig_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{V}}{4\pi ^2}D_\nu (x)[\overline{n}(x)\gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)].$$
$`(\mathrm{A}.27)`$
The effective Lagrangians Eq.(A.23) and Eq.(A.25) as well as Eq.(A.26) and Eq.(A.27) testify distinctly that the transitions p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$ and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> n + n are governed by the same dynamics of lowโenergy nuclear forces in agreement with charge independence of weak interaction strength.
For the evaluation of the effective Lagrangian $`_{\mathrm{eff}}^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(x)`$ of the transition $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p in the NNJL model one should use the following Lagrangians
$$_{\mathrm{npD}}^{}(x)=ig_\mathrm{V}[\overline{p}(x)\gamma ^\mu n^c(x)\overline{n}(x)\gamma ^\mu p^c(x)]D_\mu (x),$$
$$_{\mathrm{eff}}^{\mathrm{np}\mathrm{np}}(x)=C_{\mathrm{NN}}\{[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\mu \gamma ^5p(x)]+[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]\},$$
$$_{\mathrm{NNZ}}(x)=g_\mathrm{A}[\overline{p}(x)\gamma ^\nu \gamma ^5p(x)\overline{n}(x)\gamma ^\nu \gamma ^5n(x)]Z_\nu (x),$$
$$Z_\nu (x)\frac{G_\mathrm{F}}{2\sqrt{2}}[\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)].$$
$`(\mathrm{A}.28)`$
The effective Lagrangian $`_{\mathrm{eff}}^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(x)`$ of the lowโenergy transition $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p is then defined by
$$_{\mathrm{eff}}^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(x)=g_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{F}}{\sqrt{2}}\frac{3g_\mathrm{V}}{8\pi ^2}D_{\mu \nu }(x)[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _{\nu _\mathrm{e}}(x)].$$
$`(\mathrm{A}.29)`$
In the lowโenergy limit when $`D_{\mu \nu }(x)[\overline{p}(x)\gamma ^\mu \gamma ^5n^c(x)]2iM_\mathrm{N}D_\nu (x)[\overline{p}(x)\gamma ^5n^c(x)]`$ the effective Lagrangian reduces itself to the from
$$_{\mathrm{eff}}^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(x)=ig_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{F}}{\sqrt{2}}\frac{3g_\mathrm{V}}{4\pi ^2}D_\nu (x)[\overline{p}(x)\gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _{\nu _\mathrm{e}}(x)].$$
$`(\mathrm{A}.30)`$
For the evaluation of the matrix element of the transition $`\nu _\mathrm{e}`$ \+ D $``$ $`\nu _\mathrm{e}`$ \+ n + p one would use the wave function of the np pair in the standard form $`|n(p_1)p(p_2)=a_\mathrm{n}^{}(p_1,\sigma _1)a_\mathrm{p}^{}(p_2,\sigma _2)|0`$. The contribution of lowโenergy nuclear forces to the relative movement of the np pair in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state should be described by the infinite series of oneโnucleon bubbles evaluated at leading order in the large $`N_C`$ expansion. The result should be expressed in terms of the Sโwave scattering length $`a_{\mathrm{np}}`$ and the effective range $`r_{\mathrm{np}}`$ of lowโenergy elastic np scattering in the $`{}_{}{}^{1}\mathrm{S}_{0}^{}`$ state.
Now let us obtain the contribution of the nucleon tensor current Eq.(4.10). In terms of the structure functions the effective Lagrangian $`\delta _{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$
$$d^4x\delta _{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=$$
$$=g_\mathrm{A}\frac{C_{\mathrm{NN}}}{8\pi ^2}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{g_\mathrm{T}}{2M_\mathrm{N}}d^4x_1\frac{d^4x_2d^4k_2}{(2\pi )^4}\frac{d^4x_3d^4k_3}{(2\pi )^4}e^{ik_2\left(x_2x_1\right)}e^{ik_3\left(x_3x_1\right)}$$
$$\times \mathrm{T}([\overline{p^c}(x_1)\gamma _\alpha \gamma ^5p(x_1)]D_{\mu \nu }^{}(x_2)[\overline{\psi }_{\nu _\mathrm{e}}(x_3)\gamma _\lambda (1\gamma ^5)\psi _\mathrm{e}(x_3)])J^{\alpha \mu \nu \lambda }(k_2,k_3;Q)$$
$$g_\mathrm{A}\frac{C_{\mathrm{NN}}}{8\pi ^2}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{g_\mathrm{T}}{2M_\mathrm{N}}d^4x_1\frac{d^4x_2d^4k_2}{(2\pi )^4}\frac{d^4x_3d^4k_3}{(2\pi )^4}e^{ik_2\left(x_2x_1\right)}e^{ik_3\left(x_3x_1\right)}$$
$$\times \mathrm{T}([\overline{p^c}(x_1)\gamma ^5p(x_1)]D_{\mu \nu }^{}(x_2)[\overline{\psi }_{\nu _\mathrm{e}}(x_3)\gamma _\lambda (1\gamma ^5)\psi _\mathrm{e}(x_3)])J^{\mu \nu \lambda }(k_2,k_3;Q).$$
$`(\mathrm{A}.31)`$
The structure functions are given by
$$๐ฅ^{\alpha \mu \nu \lambda }(k_2,k_3;Q)=$$
$$=\frac{d^4k}{\pi ^2i}\mathrm{tr}\left\{\gamma ^\alpha \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}+\widehat{k}_2}\sigma ^{\mu \nu }\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\gamma ^\lambda \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}\widehat{k}_3}\right\},$$
$$๐ฅ^{\mu \nu \lambda }(k_2,k_3;Q)=$$
$$=\frac{d^4k}{\pi ^2i}\mathrm{tr}\left\{\gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}+\widehat{k}_2}\sigma ^{\mu \nu }\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}}\gamma ^\lambda \gamma ^5\frac{1}{M_\mathrm{N}\widehat{k}\widehat{Q}\widehat{k}_3}\right\},$$
$`(\mathrm{A}.32)`$
where a 4โvector $`Q=ak_2+bk_3`$ is an arbitrary shift of a virtual momentum with arbitrary parameters $`a`$ and $`b`$.
The evaluation of the structure functions Eq.(A.32) at leading order in the large $`N_C`$ expansion gives the following effective Lagrangian $`\delta _{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ caused by the contribution of the nucleon tensor current
$$\delta _{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=g_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{g_\mathrm{T}}{2\pi ^2}\left\{D_{\mu \nu }^{}(x)[\overline{p^c}(x)\gamma ^\mu \gamma ^5p(x)]\frac{ia}{2M_\mathrm{N}}^\mu D_{\mu \nu }^{}(x)[\overline{p^c}(x)\gamma ^5p(x)]\right\}$$
$$\times [\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)].$$
$`(\mathrm{A}.33)`$
It is seen that the coupling constants of the effective Lagrangian depend on the arbitrary parameter $`a`$ caused by a shift of a virtual momentum.
The analogous expression one can get for the effective Lagrangian of the transition $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n caused by the contribution of the nucleon tensor current as well
$$\delta _{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)=g_\mathrm{A}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{g_\mathrm{T}}{2\pi ^2}\left\{D_{\mu \nu }(x)[\overline{n}(x)\gamma ^\mu \gamma ^5n^c(x)]\frac{ia}{2M_\mathrm{N}}^\mu D_{\mu \nu }(x)[\overline{n}(x)\gamma ^5n^c(x)]\right\}$$
$$\times [\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)].$$
$`(\mathrm{A}.34)`$
In the lowโenergy limit when
$$D_{\mu \nu }^{}(x)[\overline{p^c}(x)\gamma ^\mu \gamma ^5p(x)]\mathrm{\hspace{0.17em}2}iM_\mathrm{N}D_\nu ^{}(x)[\overline{p^c}(x)\gamma ^5p(x)],$$
$$D_{\mu \nu }(x)[\overline{n}(x)\gamma ^\mu \gamma ^5n^c(x)]\mathrm{\hspace{0.17em}2}iM_\mathrm{N}D_\nu (x)[\overline{n}(x)\gamma ^5n(x)]$$
$`(\mathrm{A}.35)`$
the effective Lagrangians Eq.(A.33) and Eq.(A.34) can be recast into the form
$$\delta _{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=ig_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{T}}{4\pi ^2}\xi D_\nu ^{}(x)[\overline{p^c}(x)\gamma ^5p(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)],$$
$$\delta _{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)=ig_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{T}}{4\pi ^2}\xi D_\nu (x)[\overline{n}(x)\gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)],$$
$`(\mathrm{A}.36)`$
where $`\xi `$ is an arbitrary parameter related to the parameter $`a`$ as follows
$$\xi =\frac{1}{3}(4a).$$
$`(\mathrm{A}.37)`$
The total effective Lagrangians of the transitions p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$, $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p are defined by
$$_{\mathrm{eff},\mathrm{tc}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)=i(1+\overline{\xi })g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{V}}{4\pi ^2}D_\nu ^{}(x)[\overline{p^c}(x)\gamma ^5p(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)],$$
$$_{\mathrm{eff}}^{\overline{\nu }_\mathrm{e}\mathrm{D}\mathrm{e}^+\mathrm{nn}}(x)=i(1+\overline{\xi })g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{V}}{\sqrt{2}}\frac{3g_\mathrm{V}}{4\pi ^2}D_\nu (x)[\overline{n}(x)\gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _\mathrm{e}(x)],$$
$$_{\mathrm{eff}}^{\nu _\mathrm{e}\mathrm{D}\nu _\mathrm{e}\mathrm{np}}(x)=i(1+\overline{\xi })g_\mathrm{A}M_\mathrm{N}C_{\mathrm{NN}}\frac{G_\mathrm{F}}{\sqrt{2}}\frac{3g_\mathrm{V}}{4\pi ^2}D_\nu (x)[\overline{p}(x)\gamma ^5n^c(x)][\overline{\psi }_{\nu _\mathrm{e}}(x)\gamma ^\nu (1\gamma ^5)\psi _{\nu _\mathrm{e}}(x)],$$
$`(\mathrm{A}.38)`$
where $`\overline{\xi }`$ is obtained by using the relation $`g_\mathrm{T}=\sqrt{3/8}g_\mathrm{V}`$ and is defined by
$$\overline{\xi }=\sqrt{\frac{3}{8}}\xi .$$
$`(\mathrm{A}.39)`$
Under the assumption of isotopical invariance of lowโenergy nulcear forces the best agreement with the recommended value for the astrophysical factor for the solar proton burning and the contemporary experimental data on the cross sections for the antiโneutrino disintegration of the deuteron $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ e<sup>+</sup> \+ n + n and $`\overline{\nu }_\mathrm{e}`$ \+ D $``$ $`\overline{\nu }_\mathrm{e}`$ \+ n + p caused by charged and neutral weak current, respectively, we obtain at $`\overline{\xi }=0`$.
## Figure caption
* Fig.1. Oneโnucleon loop diagrams of the contribution of the effective coupling $`[\overline{p}(x)\gamma ^5n^c(x)][\overline{n^c}(x)\gamma ^5p(x)]`$ to the effective Lagrangian of the M1 transition n + p $``$ D + $`\gamma `$.
* Fig.2. Oneโnucleon loop diagrams of the contribution of the effective coupling $`[\overline{p}(x)\gamma ^\alpha \gamma ^5n^c(x)][\overline{n^c}(x)\gamma _\alpha \gamma ^5p(x)]`$ to the effective Lagrangian of the M1 transition n + p $``$ D + $`\gamma `$.
* Fig.3. Oneโnucleon loop diagrams describing the effective Lagrangian $`_{\mathrm{eff}}^{\mathrm{pp}\mathrm{De}^+\nu _\mathrm{e}}(x)`$ of the lowโenergy transition p + p $``$ D + e<sup>+</sup> \+ $`\nu _\mathrm{e}`$.
|
warning/0006/cond-mat0006431.html
|
ar5iv
|
text
|
# Systematic and Causal Corrections to the Coherent Potential Approximation
## I Introduction
The Coherent Potential Approximation (CPA) is a widely used method for treating disordered systems. Within the CPA, the problem is first averaged over all possible disorder configurations in an attempt to regain the translational invariance lost due to disorder. Then non-local correlations of the disorder potential are neglected leading to a self consistent single-site approximation (like the Weiss mean field theory of magnetism). It has been applied with great success to a variety of problems, including ab-initio calculations in disordered metallic alloys.
Nevertheless, the CPA fails to provide a completely satisfactory theory for disordered systems. As a single-site mean field theory, it cannot account for the disorder induced, short ranged but nonlocal, correlations due to the local order in the environment of each site responsible for band tailing and sharp structures in the density of states. There have been many attempts to formulate non-local corrections to the CPA in which the lattice is mapped onto a self-consistently embedded finite-sized cluster. However, as argued clearly and in detail by Gonis , these theories all fail in some significant way. A successful theory must be able to account for fluctuations in the local environment in a self-consistent way, become exact in the limit of large cluster sizes, and recover the CPA when the cluster size equals one. It must be easily implementable numerically and preserve the translational and point-group symmetries of the lattice. Finally, and most significantly, it should be fully causal so that the single-particle Green function and self energy are analytic in the upper half plane. No presently existing cluster extension of the CPA satisfies all these requirements.
Recently, a new method called the The Dynamical Cluster Approximation (DCA) was developed for ordered correlated systems such as the Hubbard model to add non-local corrections to the Dynamical Mean Field Approximation. In this manuscript, we modify the DCA to include disorder and show that the resulting formalism satisfies all of the above stated requirements for a successful cluster extension of the CPA.
In the next Section, we review the basic diagrammatic perturbation theory formalism for disordered systems. In Sec. III A we show that the CPA is equivalent to neglecting momentum conservation at all internal vertices of the diagrams, and in Sec. III B we introduce the DCA for disordered systems which systematically restores the momentum conservation relinquished by the CPA. In Sec. IV we show that the DCA satisfies each of the desired characteristics described above. In Sec. V we show results for the two-dimensional Anderson model with binary diagonal disorder. Finally in Appendix A we present an alternate way of viewing and justifying the disordered DCA algorithm developed in this paper, using the replica (or other) methods of disorder averaging.
## II Basic Formalism
We consider an Anderson model with diagonal disorder, described by the Hamiltonian
$$H=\underset{<ij>,\sigma }{}t\left(C_{i,\sigma }^{}C_{j,\sigma }+C_{j,\sigma }^{}C_{i,\sigma }\right)+\underset{i\sigma }{}(V_i\mu )n_{i,\sigma }$$
(1)
where $`C_{i,\sigma }^{}`$ creates a quasiparticle on site $`i`$ with spin $`\sigma `$, $`n_{i,\sigma }=C_{i,\sigma }^{}C_{i,\sigma }`$. The disorder occurs in the local orbital energies $`V_i`$, which we assume are independent quenched random variables distributed according to some specified probability distribution $`P(V)`$. The DCA formalism that we develop in this paper is a general method valid for any $`P(V)`$. However, for illustrative purposes, for the specific calculations presented in this paper we take $`V_i=\pm V`$ with equal probability $`1/2`$ (binary disorder).
The effect of the disorder potential $`_{i\sigma }V_in_{i,\sigma }`$ can be described using standard diagrammatic perturbation theory (although we will eventually sum to all orders). It may be re-written in reciprocal space as
$$H_{dis}=\frac{1}{N}\underset{i,๐ค,๐ค^{},\sigma }{}V_iC_{๐ค,\sigma }^{}C_{๐ค^{},\sigma }e^{i๐ซ_i(๐ค๐ค^{})}$$
(2)
The corresponding irreducible (skeletal) contributions to the self energy may be represented diagrammatically and are displayed in Fig. 1.
Here each $``$ represents the scattering of an electronic Bloch state from a local disorder potential at some site represented by $`X`$. The dashed lines connect scattering events that involve the same local potential. In each graph, the sums over the sites are restricted so that the different $`X`$ โs represent scattering from different sites. No graphs representing a single scattering event are included since these may simply be absorbed as a renormalization of the chemical potential $`\mu `$.
Translational invariance and momentum conservation are restored by averaging over all possible values of the disorder potentials $`V_i`$. For example, consider the second diagram in Fig. 1, given by
$$\frac{1}{N^3}\underset{i,๐ค_3,๐ค_4}{}V_i^3G(๐ค_3)G(๐ค_4)e^{i๐ซ_i(๐ค_1๐ค_3+๐ค_3๐ค_4+๐ค_4๐ค_2)}$$
(3)
where $`G(๐ค)`$ is the disorder-averaged single-particle Green function for state $`๐ค`$. The average over the distribution of scattering potentials $`V_i^3=V^3`$ independent of i. After summation over the remaining labels, this becomes
$$V^3G(๐ซ=0)^2\delta _{๐ค_1,๐ค_2}$$
(4)
where $`G(๐ซ=0)`$ is the local Green function. Thus the second diagramโs contribution to the self energy involves only local correlations. Since the internal momentum labels always cancel in the exponential, the same is true for all non-crossing diagrams shown in the top half of Fig. 1.
Only the diagrams with crossing dashed lines are non-local. Consider the fourth-order diagrams such as those shown on the bottom left and upper right of Fig. 1. When we impurity average, we generate potential terms $`V^4`$ when the scattering occurs from the same local potential (i.e. the third diagram) or $`V^2^2`$ when the scattering occurs from different sites, as in the fourth diagram. When the latter diagram is evaluated, to avoid overcounting, we need to subtract a term proportional to $`V^2^2`$ but corresponding to scattering from the same site. This term is needed to account for the fact that the fourth diagram should really only be evaluated for sites $`ij`$! For example, the fourth diagram yields
$`{\displaystyle \frac{1}{N^4}}{\displaystyle \underset{ij๐ค_3๐ค_4๐ค_5}{}}V_i^2V_j^2e^{i๐ซ_i(๐ค_1+๐ค_4๐ค_5๐ค_3)}e^{i๐ซ_j(๐ค_5+๐ค_3๐ค_4๐ค_2)}`$ (5)
$`G(๐ค_5)G(๐ค_4)G(๐ค_3)`$ (6)
Evaluating the disorder average $``$, we get the following two terms:
$`{\displaystyle \frac{1}{N^4}}{\displaystyle \underset{ij๐ค_3๐ค_4๐ค_5}{}}V^2^2e^{i๐ซ_i(๐ค_1+๐ค_4๐ค_5๐ค_3)}e^{i๐ซ_j(๐ค_5+๐ค_3๐ค_4๐ค_2)}`$ (7)
$`G(๐ค_5)G(๐ค_4)G(๐ค_3)`$ (8)
$`{\displaystyle \frac{1}{N^4}}{\displaystyle \underset{i๐ค_3๐ค_4๐ค_5}{}}V^2^2e^{i๐ซ_i(๐ค_1๐ค_2)}G(๐ค_5)G(๐ค_4)G(๐ค_3)`$ (9)
Momentum conservation is restored by the sum over $`i`$ and $`j`$; i.e. over all possible locations of the two scatterers. It is reflected by the Laue functions, $`\mathrm{\Delta }=N\delta _{๐ค+\mathrm{}}`$, within the sums
$`{\displaystyle \frac{\delta _{๐ค_2,๐ค_1}}{N^3}}{\displaystyle \underset{๐ค_3๐ค_4๐ค_5}{}}V^2^2N\delta _{๐ค_2+๐ค_4,๐ค_5+๐ค_3}`$ (10)
$`G(๐ค_5)G(๐ค_4)G(๐ค_3)`$ (11)
$`{\displaystyle \frac{\delta _{๐ค_2,๐ค_1}}{N^3}}{\displaystyle \underset{๐ค_3๐ค_4๐ค_5}{}}V^2^2G(๐ค_5)G(๐ค_4)G(๐ค_3)`$ (12)
Since the first term in Eq. 12 involves convolutions of $`G(๐ค)`$ it reflects non-local correlations. \[Local contributions such as the second term in Eq. 12 can, if one so chooses, be combined together with the contributions from the corresponding local diagrams such as the third diagram in Fig. 1 by replacing $`V^4`$ in the latter by the cumulant $`V^4V^2^2`$ .\] Given the fact that different $`X`$โs must correspond to different sites, it is easy to see that all crossing diagrams must involve non-local correlations.
## III Cluster Approximations
### A The Coherent Potential Approximation
In the Coherent Potential Approximation (CPA), nonlocal correlations involving different scatterers are ignored. Thus, in the calculation of the self energy, we ignore all of the crossing diagrams shown on the bottom of Fig. 1; and retain only the class of diagrams such as those shown on the top representing scattering from a single local disorder potential. These diagrams are shown in Fig. 2.
Employing diagrammatic arguments, it is easy to see that the CPA is fully equivalent to the neglect of momentum conservation at each internal vertex. This is accomplished by setting each Laue function within the sum (eg., in Eq. 12) to 1. We may then freely sum over the internal momenta, leaving only local propagators. All non-local self energy contributions (crossing diagrams) must then vanish. For example, consider again the fourth graph. If we replace the Laue function $`N\delta _{๐ค_1+๐ค_4,๐ค_5+๐ค_3}1`$ in Eq. 12, then the two contributions cancel and this diagram vanishes. Thus an alternate definition of the CPA, in terms of the Laue functions $`\mathrm{\Delta }`$, is
$$\mathrm{\Delta }=\mathrm{\Delta }_{CPA}=1$$
(13)
I.e., the CPA is equivalent to the neglect of momentum conservation at all internal vertices of the disorder-averaged irreducible graphs.
### B The Dynamical Cluster Approximation
The DCA systematically restores the momentum conservation at internal vertices which was relinquished by the CPA, and so incorporates non-local corrections. This is done by dividing the Brillouin zone into $`N_c`$ equal cells, as shown in Fig. 3 and requiring that momentum be partially conserved for momentum transfers between the coarse graining cells shown in Fig. 3, but ignored for momentum transfers within each cell. This may be accomplished by employing the Laue functions
$$\mathrm{\Delta }=\mathrm{\Delta }_{DCA}=N_c\delta _{๐(๐ค_1)+๐(๐ค_2),๐(๐ค_3)+๐(๐ค_4)\mathrm{}}$$
(14)
where $`๐(๐ค)`$, illustrated in Fig 3, maps the momenta $`๐ค`$ to the nearest cluster momenta $`๐`$. $`\mathrm{\Delta }_{DCA}`$ becomes one when $`N_c=1`$ since then all momenta are mapped to the zone center by $`๐`$. Thus the CPA is recovered in this limit. Furthermore, as $`N_c`$ becomes large, the exact result is recovered since $`lim_{N_c\mathrm{}}๐(๐ค)=๐ค`$ for all momenta $`๐ค`$.
If we employ the DCA Laue function in each of the self energy diagrams shown in Fig. 1 then we may freely sum over the momenta within each coarse-graining cell shown in Fig. 3. As illustrated in Fig. 4 for a fourth-order graph, this leads to the replacement of the lattice propagators $`G(๐ค_1)`$, $`G(๐ค_2)`$, โฆ by coarse grained propagators $`\overline{G}(๐)`$, $`\overline{G}(๐^{})`$, โฆ. which are given by
$$\overline{G}(๐)\frac{N_c}{N}\underset{\stackrel{~}{๐ค}}{}G(๐+\stackrel{~}{๐ค}),$$
(15)
where $`N`$ is the number of points of the lattice, $`N_c`$ is the number of cells in the cluster, and the $`\stackrel{~}{๐ค}`$ summation runs over the momenta of the cell about the cluster momentum $`๐`$ (cf. Fig. 3).
As $`N_c`$ increases from one, systematic non-local corrections to the CPA self energy are introduced. To see this, recall that the self energy is a functional of the Green function. According to Nyquistโs sampling theorem, to reproduce correlations of length $`L/2`$ in the Green function and corresponding self energy, we only need to sample the reciprocal space at intervals of $`\mathrm{\Delta }k2\pi /L`$. Knowledge of these Green functions on a finer scale in momentum is than $`\mathrm{\Delta }k`$ unnecessary, and may be discarded to reduce the complexity of the problem. Thus the cluster self energy will be constructed from the coarse-grained average of the single-particle Green function within the cell centered on the cluster momenta. For short distances $`rL/2`$, where $`L`$ is now the linear size of the cluster, the Fourier transform of the Green function $`\overline{G}(r)G(r)+๐ช((r\mathrm{\Delta }k)^2)`$, so that short ranged correlations are reflected in the irreducible quantities constructed from $`\overline{G}`$; whereas, longer ranged correlations $`r>L/2`$ are cut off by the finite size of the cluster .
We show in the appendix that that free-energy arguments presented previously apply to the disordered case as well. In particular, the DCA estimate of the lattice free energy is minimized by the choice $`\mathrm{\Sigma }(๐ค,\omega )=\overline{\mathrm{\Sigma }}(๐(๐ค),\omega )`$.
#### Algorithm
With the substitution $`\mathrm{\Delta }\mathrm{\Delta }_{DCA}`$, most of the diagrams represented in Fig. 1 remain. However, the complexity of the problem is greatly reduced since the nontrivial sums involve only the cluster momenta $`K`$ ( numbering $`N_c`$ instead of $`N`$). Furthermore, since these diagrams are the same as those from a finite-sized periodic cluster of $`N_c`$ sites, we can easily sum this series to all orders by numerically solving the corresponding cluster problem. The resulting algorithm is as follows (for notational convenience the frequency arguments are not displayed below):
1. Make a guess for the impurity-averaged cluster self energy $`\overline{\mathrm{\Sigma }}(๐)`$, usually zero.
2. Calculate the coarse-grained Cluster Green functions
$$\overline{G}(๐)=\frac{N_c}{N}\underset{\stackrel{~}{๐ค}}{}\frac{1}{\omega +\mu ฯต_{๐+\stackrel{~}{๐ค}}\overline{\mathrm{\Sigma }}(๐)}$$
(16)
3. Calculate the cluster-excluded propagator $`๐ข(๐)=1/(1/\overline{G}(๐)+\overline{\mathrm{\Sigma }}(๐))`$. The introduction of $`๐ข(๐)`$ is necessary to avoid overcounting diagrams on the cluster.
4. Fourier transform $`๐ข`$ to the real space matrix representation of the cluster problem, (i.e., write $`๐ข_{n,m}=_๐๐ข(๐)\mathrm{exp}i๐(๐ซ_n๐ซ_m)`$ ) whence the disorder potential may be represented as a diagonal matrix $`๐`$ (with elements $`V_n`$ on the cluster sites labelled by $`n`$) and form a new estimate of the disorder-averaged cluster Green function matrix
$$๐=\left(๐ข^1๐\right)^1$$
(17)
where the average $``$ indicates an average over disorder configurations on the cluster .
5. Transform back to the cluster reciprocal space and form a new estimate of the self energy $`\overline{\mathrm{\Sigma }}(๐)=1/๐ข(๐)1/G(๐)`$
6. Repeat, starting from 2, until $`\overline{\mathrm{\Sigma }}(๐)`$ converges to the desired accuracy.
The algorithm recovers the CPA for $`N_c=1`$ and becomes exact when $`N_c\mathrm{}`$.
## IV Characteristics of the DCA
In Ref. A. Gonis discusses the CPA, and various methods to incorporate non-local corrections. He lists the most important characteristics of a successful cluster theory. A successful theory must be able to account for fluctuations in the local environment in a self-consistent way, become exact in the limit of large cluster sizes, and recover the CPA when the cluster size becomes one. It must be straight-forward to implement numerically and preserve the full point-group symmetry of the lattice. Finally, and most significantly, it should be fully causal so that the single-particle Green function and self energy are analytic in the upper half plane. In the next two sections, we demonstrate that the DCA satisfies each of these requirements.
#### The limits $`N_c1`$ and $`N_c\mathrm{}`$
: As mentioned above, the DCA recovers the CPA for $`N_c=1`$. When $`N_c=1`$, $`๐=0`$, and $`\stackrel{~}{๐ค}=๐ค`$. Then the DCA algorithm reduces to the self consistent scalar equations (in contrast to the DCA which involves matrix equations):
$$\overline{G}=\frac{1}{N}\underset{๐ค}{}\frac{1}{\omega +\mu ฯต_๐ค\overline{\mathrm{\Sigma }}},$$
(18)
$$๐ข^1=1/\overline{G}+\overline{\mathrm{\Sigma }},$$
(19)
$$\overline{G}=\left(๐ข^1V\right)^1,$$
(20)
which together correspond exactly to the prescription for CPA , and are easily solved, for example by iteration.
As $`N_c\mathrm{}`$ the DCA becomes exact, including correlations over all length scales. For, the DCA Laue function requires complete momentum conservation in this limit, i.e., $`[๐,\overline{G}(๐)][๐ค,G(๐ค)]`$, whence $`๐ข=1/(\omega +\mu ฯต_๐ค)`$ which is the bare lattice Green function so that Eq. 12 amounts to solving the problem by exact diagonalization.
#### Causality
: It is easy to show that this algorithm is fully causal.
For our purposes, this is equivalent to requiring that all the retarded propagators and self energies are Herglotz, or analytic in the upper half plane. Since the diagrams which describe the DCA cluster problem are isomorphic to those of a real finite-size periodic cluster, the corresponding impurity averaged cluster self energy shares the causal properties of this system. Furthermore, the coarse graining step (2) cannot violate causality, since the sum of analytic functions is analytic. The only โsuspectโ step is the cluster-exclusion step (3), thus we must show that the imaginary part of $`๐ข`$ is negative semidefinite. The essential steps of the argument are sketched in Fig. 5. The imaginary part of $`๐ข(๐,\omega )=(\overline{G}(๐,\omega )^1+\overline{\mathrm{\Sigma }}(๐,\omega ))^1`$ is negative provided that $`\mathrm{Im}(\overline{G}(๐,\omega )^1)\mathrm{Im}\overline{\mathrm{\Sigma }}(๐,\omega )`$. $`\overline{G}(๐,\omega )`$ can be written as $`\overline{G}(๐,\omega )=(N_c/N)_{\stackrel{~}{๐ค}}(z_{๐+\stackrel{~}{๐ค}})^1(\omega )`$, where the $`z_{๐+\stackrel{~}{๐ค}}(\omega )`$ are complex numbers with a positive semidefinite imaginary part $`\text{Im}\overline{\mathrm{\Sigma }}(๐,\omega )`$. For any $`๐`$ and $`\omega `$, the set of points $`z_{๐+\stackrel{~}{๐ค}}(\omega )`$ are on a segment of the dashed horizontal line in the upper half plane due to the fact that the imaginary part is independent of $`\stackrel{~}{๐ค}`$. The mapping $`z1/z`$ maps this line segment onto a segment of the dashed circle shown in the lower half plane. $`\overline{G}(๐,\omega )`$ is obtained by summing the points on the circle segment, yielding the empty dot that must lie within the dashed circle. The inverse necessary to take $`\overline{G}(๐,\omega )`$ to $`1/\overline{G}(๐,\omega )`$ maps this point onto the empty dot in the upper half plane which must lie above the dashed line. Thus, the imaginary part of $`\overline{G}(๐,\omega )^1`$ is greater than or equal to $`\mathrm{Im}\overline{\mathrm{\Sigma }}(๐,\omega )`$. This argument may easily be extended for $`๐ข(z)`$ for any $`z`$ in the upper half plane. Thus $`๐ข`$ is completely analytic in the upper half plane.
#### Preservation of the Lattice Symmetry
: Since the DCA is formulated in reciprocal space, it preserves the translational symmetry of the system. However, care must be taken when selecting the coarse-graining cells to preserve the point group symmetries of the lattice. For example, the calculations presented in Sec. V are done for a simple square lattice. Both it and its reciprocal lattice have a $`C_{4v}`$ symmetry with eight point group operations. We must choose a set of coarse-graining cells which preserve this point-group symmetry. This may be done by tiling the real lattice with squares, and using the $`๐`$ points that correspond to the reciprocal space of the tiling centers. For large $`V`$, an important configuration of the half filled system is that with all of the sites with $`V_i=V`$ occupied and those with $`V_i=+V`$ unoccupied. To retain this configuration on the cluster, when $`N_c>1`$ we will choose $`N_c`$ even.
Square tilings with an even number of sites include $`N_c=4,8,10,16,18,20,26,32,34,36,\mathrm{}`$. The first few are illustrated in Fig. 6. The relation between the principal lattice vectors of the lattice centers, $`๐_1`$ and $`๐_2`$, and the reciprocal lattice takes the usual form $`๐ _i=2\pi ๐_i/\left|๐_1\times ๐_2\right|`$, with $`๐_{nm}=n๐ _1+m๐ _2`$ for integer $`n`$ and $`m`$. For tilings with either $`a_{1x}=a_{1y}`$ (corresponding to $`N_c=8,18,32\mathrm{}`$) or one of $`a_{1x}`$ or $`a_{1y}`$ zero (corresponding to $`N_c=1,4,16,36\mathrm{}`$), the principal reciprocal lattice vectors of the coarse-grained system either point along the same directions as the principal reciprocal lattice vectors of the real system or are rotated from them by $`\pi /4`$. As a result equivalent momenta $`๐ค`$ are always mapped to equivalent coarse-grained momenta $`๐`$. An example for $`N_c=8`$ is shown in Fig. 7. However, for $`N_c=10,20,26,34\mathrm{}`$, the principal reciprocal lattice vectors of the coarse-grained system do not point along a high symmetry direction of the real lattice. Since all points within a coarse-graining cell are mapped to its center $`๐`$, this means that these coarse graining choices violate the point group symmetry of the real system. This is illustrate for $`N_c=10`$ in Fig. 7, where the two open dots, resting at equivalent points in the real lattice, fall in inequivalent coarse graining cells and so are mapped to inequivalent $`๐`$ points. Thus the tilings corresponding to $`N_c=10,20,26,34\mathrm{}`$ violate the point-group symmetry of the real lattice system and should be avoided.
#### An efficient numerical algorithm for disorder averaging
: The implementation of the DCA clearly requires an efficient algorithm for disorder averaging on a cluster of size $`N_c`$.( Needless to say, this particular aspect is common to all approaches where disorder averaging is involved.) Even for a system with binary diagonal disorder, where each site can acquire only one of two values for the potential ($`\pm V`$) the total number of disorder configurations is $`2^{N_c}`$, which grows exponentially with $`N_c`$. For a generic quenched disorder, the probability of the various disorder configurations are determined by the specified $`P(V)`$; whereas, for annealed disorder, the probability of a configuration depends upon an effective Boltzmann factor determined by the electronic partition function for that particular disorder configuration .
In this section, we propose an approach to carrying out the disorder averaging by statistically sampling disorder configurations using a Markov process. For systems with quenched disorder, one could also sample random configurations of the disorder potentials and calculate the corresponding Green function using matrix inversion. A significant advantage of the Markov technique is that it may be easily modified to treat either quenched or annealed disorder.
In a Markov process, each disorder configuration depends upon the previous configurations. To evolve from one configuration to another, we will propose local changes in the disorder potentials. These changes are accepted with some probability determined by either the effective Boltzmann factor , for annealed disorder, or the probability distribution $`P(V_i)`$ , for quenched disorder. If we accept such a change in one of the disorder potentials, say on site $`l`$, then the new Green function matrix $`G_{n,m}^{}`$ depends on the previous Green function matrix $`G_{n,m}`$ through the matrix relationship (where, once again, for notational convenience the frequency arguments are not displayed)
$$๐_{}^{}{}_{}{}^{1}๐^1=๐๐^{}$$
(21)
Since we change on the potential on the site $`l`$ and the matrices $`๐`$ and $`๐^{}`$ are diagonal, their difference $`\delta ๐=๐๐^{}`$ is a diagonal matrix with only the $`l`$โth diagonal element finite. Then
$$G_{n,m}^{}=G_{n,m}+G_{n,l}\delta V_lG_{l,m}^{}$$
(22)
If we set $`l=n`$ in Eq. 22, we get
$$G_{l,m}^{}=\frac{G_{l,m}}{1G_{l,l}\delta V_l}$$
(23)
If we substitute this result back into Eq. 22, we get
$$G_{n,m}^{}=G_{n,m}+G_{n,l}\frac{\delta V_l}{1G_{l,l}\delta V_l}G_{l,m}$$
(24)
an equation which requires roughly $`N_c^2`$ operations to evaluate for each frequency.
Because the technique involves importance sampling, it is likely to miss rare configurations of disorder, and any special physics that arises from such configurations, such as Lifshitz tails in the DOS. However, when these configurations are well known, we can easily adapt this method to include them. This may be done by excluding them from the sampling, and then including the corresponding configurations in the sample with the appropriate reweighting.
## V Results
### A Single-Particle Properties
To illustrate the algorithm discussed above, we present calculations on a two-dimensional square lattice system with $`t=0.25`$ and binary random disorder so that $`V_i=\pm V`$ with equal probability. No tricks are used to force the algorithm to remain causal such as renormalizing the spectrum or cutting off any negative tails of the density of states.
The self energy is plotted in Fig. 8 for $`N_c=1`$ and for $`N_c=16`$ for different values of $`๐`$ for $`V=0.1`$ (left) and $`V=0.5`$ (right). For $`V=0.1`$ the self energy for $`N_c=16`$ has very little momentum dependence, thus the different curves fall atop of one another. They also are hence very close to the self energy for $`N_c=1`$ (the CPA result) indicating that it is a very good approximation for the self energy when $`V`$ is small. However, for larger $`V=0.5`$ the self energy curves at different values of $`K`$ for $`N_c=16`$ differ considerably from each other and from the CPA self energy obtained with $`N_c=1`$. Thus, as $`V`$ increases non-local corrections clearly become important and are expressed in the momentum-dependence of the self energy.
As shown in Fig. 9 (top left) the single particle density of states is essentially independent of $`N_c`$ for $`V=0.1`$; however, for $`V=1.0`$, the density of states depends strongly upon $`N_c`$. In particular the gap around $`\omega =0`$, which is sharp for $`N_c=1`$, is partially filled in. The top and bottom of the band also acquire tails as $`N_c`$ increases. When $`N_c>1`$, the density of states acquires several additional structures which correspond to important local configurations of the disorder. The additional features and the band tails are absent in the CPA and believed to be due to local order in the environment of each site.
### B (Absence of) Localization
Despite its advantages over the CPA as discussed above, one feature that the DCA shares with the CPA and similar self-consistent cluster methods is its limited ability to take into account localization effects . To show this, we measure the probability that an electron remains at site $`l`$ for all time:
$`P(\mathrm{})`$ $`=`$ $`\underset{t\mathrm{}}{lim}\left|G(l,l,t)\right|^2`$ (25)
$`=`$ $`\underset{\eta 0}{lim}{\displaystyle \frac{\eta }{\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐ฯต\left|G_{l,l}(ฯต+i\eta )\right|^2.`$ (26)
As shown in, $`P(\mathrm{})`$ is expected to be nonzero as long as there are a thermodynamically significant fraction of localized states in the spectrum of eigenstates of the disordered system. In 1 and 2 dimensions this is expected to happen for arbitrarily small but thermodynamically significant disorder. Since the cluster is formed by coarse-graining the real-lattice problem in reciprocal space, local quantities on the cluster and the real lattice correspond one-to-one. Thus, to test for localization, we need only apply the formula 26 for each site on the cluster. Making this substitution and introducing the local coarse-grained (but not disorder averaged) spectral function, $`\overline{A}(l,\omega )=\frac{1}{\pi }\mathrm{Im}\overline{G}_{l,l}(\omega )`$, Eq. 26 becomes
$`P(\mathrm{})`$ $`=`$ $`\underset{\eta 0}{lim}p(\eta )`$ (27)
$`=`$ $`\underset{\eta 0}{lim}{\displaystyle \frac{2i\eta }{N_c}}{\displaystyle \underset{l}{}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐\omega ๐\omega ^{}{\displaystyle \frac{\overline{A}(l,\omega )\overline{A}(l,\omega ^{})}{\omega \omega ^{}2i\eta }}.`$ (28)
$`p(\eta )`$ is plotted versus $`\eta `$ in the inset to Fig. 10 for the half-filled model when $`V=0.4`$. The $`p(\eta )`$ extrapolates to zero, indicating the lack of localization.
This result can be understood from either a diagrammatic perspective or by carefully assessing the cluster problem. As is well known , the crossing diagrams, especially those which involve many crossings, describe the coherent backscattering of electrons which are responsible for localization. Hence the CPA, which includes only non-crossing diagrams, can not describe localization. Within the DCA, however, for $`N_c>1`$ some crossing graphs are restored . Within each diagram, each $`X`$ represents scattering from distinct site. Since there are only $`N_c`$ sites on the cluster, the maximally crossed DCA graphs can have at most $`N_c`$ crossings. Since all states are expected to be localized in two dimensional disordered system, apparently an infinite number of crossings are needed to describe localization diagrammatically. From the perspective of the cluster, this result is not surprising since each site on the cluster is coupled to a non-interacting translationally invariant host into which electrons can escape. Thus, if the density of states is finite at some energy, then the corresponding states can not be localized unless the hybridization rate at that energy between the cluster and the host vanishes.
As described in Ref. , the hybridization rate between the cluster and its host is given by
$$\mathrm{\Gamma }(๐,\omega )=\mathrm{Im}\left(\frac{1}{\overline{G}(๐,\omega )}+\overline{\mathrm{\Sigma }}(๐,\omega )\right)$$
(29)
The net hybridization rate to a site on the cluster (the $`๐`$-integrated $`\mathrm{\Gamma }(๐,\omega )`$) is plotted in Fig. 10 when $`V=0.4`$ for several values of $`N_c`$. It remains finite over the entire region where the corresponding density of states, shown in Fig. 9, is finite. This is consistent with the lack of localization demonstrated in the inset.
We note that the hybridization falls as $`N_c`$ increases (for large $`N_c`$, $`\mathrm{\Gamma }(๐,\omega )๐ช(1/N_c)`$ ) especially at the band edges, although, given that it is defined entirely in terms of disorder-averaged propagators, it is still unlikely to be sensitive to localization effects. However, the number of diagrammatic crossings in two particle properties (such as the conductivity) which are strongly affected by localization effects does increase with $`N_c`$ even within the DCA. Thus, it is likely that disordered DCA can describe the precursor effects of localization. Some evidence for this can be seen in the (finite time) probability that an electron on a site $`l`$ remains after a time $`t`$, $`P(t)=\left|G(l,l,t)\right|^2`$. As shown in Fig. 11, for $`N_c=1`$, this probability falls quickly with time. The long time behavior is shown in the inset. As $`N_c`$ increases, the electron remains localized for longer times. Hence one can hope that a careful finite size scaling study of two particle properties within the disordered DCA can even capture some aspects of the localization transition.
## VI Conclusion
We have developed a modification of the Dynamical Cluster Approximation to treat disordered systems. This formalism satisfies all of the characteristics of a successful cluster approximation. It is causal, preserves the point-group and translational symmetry of the original lattice, recovers the CPA when the cluster size goes to one, and becomes exact as $`N_c\mathrm{}`$. Like the CPA the problem is disorder averaged and has a simple diagrammatic formulation. It is easy to implement numerically and restores sharp features and band tailing in the DOS which reflect correlations in the local environment of each site. Although the DCA does not capture the localization transition, it does describe the precursor effects of localization. It systematically restore the crossing graphs known to be responsible for localization, and might be able to access the localization transition itself via an appropriate finite size scaling analysis of two particle properties which remains to be developed.
The DCA formalism we have discussed here can also be extended to problems with disorder and interactions simply by incorporating interaction diagrams in the self energy. This is also discussed in the Appendix below. The DCA should be able to provide a good description of localization effects at finite temperatures in such contexts. For, in such cases the scattering processes are partially inelastic, so that the coherent back scattering disappears after a characteristic inelastic scattering time. In this time only a finite-number of back-scattering processes can occur so only a finite number of diagrammatic crossings are needed to describe the finite-temperature physics, and these are captured in the DCA.
#### Acknowledgements
This work was initiated in conversations with B.L. Gyorffy. It is a pleasure to acknowledge useful discussions with F.P. Esposito, A. Gonis, M. Hettler, D. E. Logan, and M. Ma . This work was supported in part by NSF grants DMR-9704021, DMR-9357199, and PHY94-07194 and by PRF grant ACF-PRF#33611-AC6. This research was supported in part by NSF cooperative agreement ACI-9619020 through computing resources provided by the National Partnership for Advanced Computational Infrastructure at the San Diego Supercomputer Center.
## A Disorder-DCA from the Replica Method
An alternate way of justifying the DCA in the context of disordered systems is to use the replica (or other such) trick for disorder averaging . For, this maps the disorder averaged problem into what looks like an interacting problem, whence the DCA formalism developed by us earlier can simply be transcribed for this case, to arrive at the appropriate self consistent cluster problem. For the effective cluster problem, the replica trick can be โun-doneโ, and we recover the algorithm presented earlier in this paper. The same procedure also works for problems involving both disorder and interactions. We detail this below.
As is well known, for problems involving quenched disorder, as for example corresponding to the Hamiltonian:
$`H`$ $`=`$ $`H_0+H_{dis}`$ (A1)
$`H_0`$ $`=`$ $`{\displaystyle \underset{๐ค,\sigma }{}}\xi _๐คC_{๐ค,\sigma }^{}C_{๐ค,\sigma }`$ (A2)
$`H_{dis}`$ $`=`$ $`{\displaystyle \underset{i\sigma }{}}V_in_{i,\sigma }`$ (A3)
where $`\xi _๐ค=ฯต_๐ค\mu `$, and $`V_i`$ is the random potential distributed according to a given probability distribution $`P(V)`$, complications arise because the disorder averaging has to be done on the free energy
$$F=k_BT\mathrm{ln}Z$$
and the green functions
$$G_{i,j}(\tau )=Tr[๐_\tau C_{i,\sigma }(\tau )C_{j,\sigma }^{}exp(\beta H)]/Z.$$
Here $`Z=Tr[exp(\beta H)]`$ is the partition function, and $`๐_\tau `$ represents the imaginary-time ordering operator.
In the replica trick , one writes
$$\mathrm{ln}Z=\underset{m_r0}{lim}\frac{Z^{m_r}1}{m_r}\text{ and }1/Z=\underset{m_r0}{lim}Z^{m_r1}$$
and assumes that the order of taking the limit $`m_r0`$ and disorder averaging can be interchanged. Then for any positive integer $`m_r`$, the resulting disorder averaged quantities such as $`Z^{m_r}`$ , $`G_๐ค(\tau )`$, etc., can be represented in terms of an interacting problem involving $`m_r`$ replicas of the original electronic degrees of freedom, which we index with the subscript $`\alpha =1,\mathrm{},m_r`$.
For example, using the standard Fermionic (Grassmann variable) functional integrals to represent the traces above, we can write
$`Z^{m_r}`$ $`=`$ $`{\displaystyle Dc^{}Dc\mathrm{exp}[\beta \mathrm{\Psi }]}`$ (A4)
$`G_๐ค(\tau )`$ $`=`$ $`{\displaystyle Dc^{}Dcc_{๐ค,\sigma ,1}(\tau )c_{๐ค,\sigma ,1}^{}(0)\mathrm{exp}[\beta \mathrm{\Psi }]}`$ (A5)
Here $`\mathrm{\Psi }`$ is an effective free energy functional which arises from the disorder averaging, and can be written as
$$\beta \mathrm{\Psi }=\underset{๐ค,\sigma ,\alpha }{}_0^\beta ๐\tau c_{๐ค,\sigma ,\alpha }^{}(\tau )(_\tau +\xi _๐ค)c_{๐ค,\sigma ,\alpha }(\tau )+\underset{i}{\overset{N}{}}W(\stackrel{~}{n}_i)$$
(A6)
where,
$$\stackrel{~}{n}_i\underset{\alpha ,\sigma }{}_0^\beta n_{i,\sigma ,\alpha }(\tau )๐\tau $$
and
$$\mathrm{exp}[W(\stackrel{~}{n}_i)]=\mathrm{exp}(V_i\stackrel{~}{n}_i)=๐V_iP(V_i)exp(V_i\stackrel{~}{n}_i)$$
In terms of the cumulants $`V^l_c`$ of the disorder distribution $`P(V)`$, one can write
$$W(\stackrel{~}{n}_i)=\underset{l=2}{\overset{\mathrm{}}{}}\frac{1}{l!}V^l_c(\stackrel{~}{n}_i)^l$$
So, clearly, W introduces (local in space but non-local in time) interactions between electrons belonging to arbitrary replicas.
If one re-expands $`\mathrm{exp}[W(\stackrel{~}{n}_i)]`$ in powers of the cumulants $`V^l_c`$ one can perform the Fermionic traces using the standard techniques of diagrammatic perturbation theory. Then, order-by-order in perturbation theory, the dependence on $`m_r`$ is explicit and analytic, and the $`lim_{m_r0}`$ can be evaluated precisely. The resulting terms are in exact, one-to-one correspondence with the terms obtainable by writing out the diagrams from a direct perturbation expansion in powers of $`V_i`$ and then disorder averaging as discussed in Section II. The $`m_r0`$ limit eliminates the diagrams (in the interacting problem ) containing internal loops with free sums over the replica indices (as required, since such diagrams never appear in the direct disorder averaged perturbation theory formalism of Section II ).
For the โreplicated interacting problemโ obtained above, one can transcribe exactly the DCA formalism discussed in refs. . If one assumes that the self-consistent host propagators do not break replica symmetry, then the effective cluster problem corresponds to a Fermionic functional integral involving an effective, self consistent cluster free energy functional given by
$`\beta \mathrm{\Psi }_c`$ $`=`$ $`{\displaystyle \underset{๐,\sigma ,\alpha }{}}{\displaystyle _0^\beta }๐\tau {\displaystyle _0^\beta }๐\tau ^{}c_{๐,\sigma ,\alpha }^{}(\tau )๐ข^1(๐,\tau \tau ^{})c_{๐,\sigma ,\alpha }(\tau ^{})`$ (A8)
$`+{\displaystyle \underset{i}{\overset{N_c}{}}}W(\stackrel{~}{n}_i)`$
But, as is easy to see using the same procedure as outlined earlier in this appendix, such an effective free energy functional is exactly what one would obtain if one were to disorder average (using the replica trick) a cluster problem with $`N_c`$ sites which are dual to the cluster momenta $`๐`$, a bare ( retarded ) cluster propagator $`๐ข^\mathrm{๐}(๐,\tau \tau ^{})`$, and a random potential $`V_i`$ distributed according to $`P(V_i)`$ at every site $`i`$ of the cluster. Hence we have an alternate justification for the disorder-DCA algorithm set down in Section III. The above route also enables one to quickly extend our discussions in ref. regarding the 2-particle propagators, Ward identities, etc., to the disorder-DCA context. Most significantly, the DCA estimate of the lattice self energy is minimized by the choice $`\mathrm{\Sigma }_\alpha (๐ค,\omega )=\overline{\mathrm{\Sigma }}(๐(๐ค),\omega )`$.
We note that the arguments presented in the main text and in this appendix are also easily extended to problems involving interactions and disorder. For example, for the case of the Hubbard model with diagonal disorder, one would add to the starting Hamiltonian the interaction term $`U_i^Nn_{i,}n_{i,}`$. Going through exactly the same procedures as outlined above, it is not hard to see that the only change is that the effective free energy functionals for the lattice and the cluster pick up the additional terms $`U_\alpha _{i^N}n_{i,,\alpha }n_{i,,\alpha }`$ and $`U_\alpha _{i^{N_c}}n_{i,,\alpha }n_{i,,\alpha }`$. The resulting cluster problem now has both interactions and disorder on the cluster of $`N_c`$ sites which are dual to the cluster momenta $`๐`$: a bare ( retarded ) cluster propagator $`๐ข^\mathrm{๐}(๐,\tau \tau ^{})`$, a random potential $`V_i`$ distributed according to $`P(V_i)`$, and the Hubbard interaction U at every site $`i`$ of the cluster. One can resort to any technique of oneโs choice to solve this problem for the disorder- averaged cluster Green functions $`\overline{G}(๐,\omega )`$ and cluster self-energies $`\overline{\mathrm{\Sigma }}(๐,\omega )`$ and go through with the rest of the DCA iteration.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.